Nothing Special   »   [go: up one dir, main page]

Introduction To Anthropology 2

Download as pdf or txt
Download as pdf or txt
You are on page 1of 222

Introduction to

Anthropology

SENIOR CONTRIBUTING AUTHORS


JENNIFER HASTY, UNIVERSITY OF PENNSYLVANIA
DAVID G. LEWIS, OREGON STATE UNIVERSITY
MARJORIE M. SNIPES, UNIVERSITY OF WEST GEORGIA
OpenStax
Rice University
6100 Main Street MS-375
Houston, Texas 77005

To learn more about OpenStax, visit https://openstax.org.


Individual print copies and bulk orders can be purchased through our website.

©2022 Rice University. Textbook content produced by OpenStax is licensed under a Creative Commons
Attribution 4.0 International License (CC BY 4.0). Under this license, any user of this textbook or the textbook
contents herein must provide proper attribution as follows:

- If you redistribute this textbook in a digital format (including but not limited to PDF and HTML), then you
must retain on every page the following attribution:
“Access for free at openstax.org.”
- If you redistribute this textbook in a print format, then you must include on every physical page the
following attribution:
“Access for free at openstax.org.”
- If you redistribute part of this textbook, then you must retain in every digital format page view (including
but not limited to PDF and HTML) and on every physical printed page the following attribution:
“Access for free at openstax.org.”
- If you use this textbook as a bibliographic reference, please include
https://openstax.org/details/books/introduction-anthropology in your citation.

For questions regarding this licensing, please contact support@openstax.org.

Trademarks
The OpenStax name, OpenStax logo, OpenStax book covers, OpenStax CNX name, OpenStax CNX logo,
OpenStax Tutor name, Openstax Tutor logo, Connexions name, Connexions logo, Rice University name, and
Rice University logo are not subject to the license and may not be reproduced without the prior and express
written consent of Rice University.

HARDCOVER BOOK ISBN-13 978-1-711494-99-9


B&W PAPERBACK BOOK ISBN-13 978-1-711494-98-2
DIGITAL VERSION ISBN-13 978-1-951693-99-2
ORIGINAL PUBLICATION YEAR 2022
1 2 3 4 5 6 7 8 9 10 RS 22
OPENSTAX

OpenStax provides free, peer-reviewed, openly licensed textbooks for introductory college and Advanced
Placement® courses and low-cost, personalized courseware that helps students learn. A nonprofit ed tech
initiative based at Rice University, we’re committed to helping students access the tools they need to complete
their courses and meet their educational goals.

RICE UNIVERSITY

OpenStax, OpenStax CNX, and OpenStax Tutor are initiatives of Rice University. As a leading research university
with a distinctive commitment to undergraduate education, Rice University aspires to path-breaking research,
unsurpassed teaching, and contributions to the betterment of our world. It seeks to fulfill this mission by
cultivating a diverse community of learning and discovery that produces leaders across the spectrum of human
endeavor.

PHILANTHROPIC SUPPORT

OpenStax is grateful for the generous philanthropic partners who advance our mission to improve educational

access and learning for everyone. To see the impact of our supporter community and our most updated list of

partners, please visit openstax.org/impact.

Arnold Ventures Burt and Deedee McMurtry

Chan Zuckerberg Initiative Michelson 20MM Foundation

Chegg, Inc. National Science Foundation

Arthur and Carlyse Ciocca Charitable Foundation The Open Society Foundations

Digital Promise Jumee Yhu and David E. Park III

Ann and John Doerr Brian D. Patterson USA-International Foundation

Bill & Melinda Gates Foundation The Bill and Stephanie Sick Fund
Girard Foundation Steven L. Smith & Diana T. Go

Google Inc. Stand Together

The William and Flora Hewlett Foundation Robin and Sandy Stuart Foundation

The Hewlett-Packard Company The Stuart Family Foundation

Intel Inc. Tammy and Guillermo Treviño

Rusty and John Jaggers Valhalla Charitable Foundation

The Calvin K. Kazanjian Economics Foundation White Star Education Foundation

Charles Koch Foundation Schmidt Futures

Leon Lowenstein Foundation, Inc. William Marsh Rice University

The Maxfield Foundation


Study where you want, what
you want, when you want.
When you access your book in our web view, you can use our new online
highlighting and note-taking features to create your own study guides.

Our books are free and flexible, forever.


Get started at openstax.org/details/books/introduction-to-anthropology

Access. The future of education.


openstax.org
Contents
Preface 1

CHAPTER 1
What Is Anthropology? 7
Introduction 7
1.1 The Study of Humanity, or "Anthropology Is Vast" 8
1.2 The Four-Field Approach: Four Approaches within the Guiding Narrative 14
1.3 Overcoming Ethnocentrism 19
1.4 Western Bias in Our Assumptions about Humanity 21
1.5 Holism, Anthropology’s Distinctive Approach 25
1.6 Cross-Cultural Comparison and Cultural Relativism 28
1.7 Reaching for an Insider’s Point of View 30
Key Terms 34
Summary 34
Critical Thinking Questions 35
Bibliography 35

CHAPTER 2
Methods: Cultural and Archaeological 39
Introduction 39
2.1 Archaeological Research Methods 40
2.2 Conservation and Naturalism 44
2.3 Ethnography and Ethnology 50
2.4 Participant Observation and Interviewing 53
2.5 Quantitative and Qualitative Analysis 55
2.6 Collections 57
Key Terms 63
Summary 64
Critical Thinking Questions 65
Bibliography 65

CHAPTER 3
Culture Concept Theory: Theories of Cultural Change 67
Introduction 67
3.1 The Homeyness of Culture 68
3.2 The Winkiness of Culture 70
3.3 The Elements of Culture 73
3.4 The Aggregates of Culture 78
3.5 Modes of Cultural Analysis 82
3.6 The Paradoxes of Culture 86
Key Terms 94
Summary 95
Critical Thinking Questions 95
Bibliography 95
CHAPTER 4
Biological Evolution and Early Human Evidence 97
Introduction 97
4.1 What Is Biological Anthropology? 98
4.2 What’s in a Name? The Science of Taxonomy 101
4.3 It’s All in the Genes! The Foundation of Evolution 103
4.4 Evolution in Action: Past and Present 108
4.5 What Is a Primate? 121
4.6 Origin of and Classification of Primates 125
4.7 Our Ancient Past: The Earliest Hominins 129
Key Terms 137
Summary 139
Critical Thinking Questions 139
Bibliography 139

CHAPTER 5
The Genus Homo and the Emergence of Us 141
Introduction 141
5.1 Defining the Genus Homo 142
5.2 Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus 145
5.3 The Emergence of Us: The Archaic Homo 151
5.4 Tracking Genomes: Our Human Story Unfolds 168
Key Terms 174
Summary 174
Critical Thinking Questions 175
Bibliography 175

CHAPTER 6
Language and Communication 179
Introduction 179
6.1 The Emergence and Development of Language 180
6.2 Language and the Mind 188
6.3 Language, Community, and Culture 193
6.4 Performativity and Ritual 196
6.5 Language and Power 199
Key Terms 206
Summary 206
Critical Thinking Questions 206
Bibliography 207

CHAPTER 7
Work, Life, and Value: Economic Anthropology 211
Introduction 211
7.1 Economies: Two Ways to Study Them 212
7.2 Modes of Subsistence 213
7.3 Gathering and Hunting 214
7.4 Pastoralism 219
7.5 Plant Cultivation: Horticulture and Agriculture 224
7.6 Exchange, Value, and Consumption 232

Access for free at openstax.org


7.7 Industrialism and Postmodernity 237
Key Terms 246
Summary 247
Critical Thinking Questions 247
Bibliography 247

CHAPTER 8
Authority, Decisions, and Power: Political Anthropology 251
Introduction 251
8.1 Colonialism and the Categorization of Political Systems 252
8.2 Acephalous Societies: Bands and Tribes 254
8.3 Centralized Societies: Chiefdoms and States 257
8.4 Modern Nation-States 264
8.5 Resistance, Revolution, and Social Movements 272
Key Terms 276
Summary 277
Critical Thinking Questions 277
Bibliography 278

CHAPTER 9
Social Inequalities 281
Introduction 281
9.1 Theories of Inequity and Inequality 282
9.2 Systems of Inequality 290
9.3 Intersections of Inequality 296
9.4 Studying In: Addressing Inequities within Anthropology 301
Key Terms 304
Critical Thinking Questions 305
Bibliography 306

CHAPTER 10
The Global Impact of Human Migration 311
Introduction 311
10.1 Peopling of the World 312
10.2 Early Global Movements and Cultural Hybridity 315
10.3 Peasantry and Urbanization 322
10.4 Inequality along the Margins 324
Key Terms 334
Summary 334
Critical Thinking Questions 335
Bibliography 335

CHAPTER 11
Forming Family through Kinship 339
Introduction 339
11.1 What Is Kinship? 340
11.2 Defining Family and Household 342
11.3 Reckoning Kinship across Cultures 347
11.4 Marriage and Families across Cultures 353
Key Terms 361
Summary 362
Critical Thinking Questions 363
Bibliography 363

CHAPTER 12
Gender and Sexuality 367
Introduction 367
12.1 Sex, Gender, and Sexuality in Anthropology 368
12.2 Performing Gender Categories 376
12.3 The Power of Gender: Patriarchy and Matriarchy 383
12.4 Sexuality and Queer Anthropology 388
Key Terms 395
Summary 395
Critical Thinking Questions 396
Bibliography 396

CHAPTER 13
Religion and Culture 399
Introduction 399
13.1 What Is Religion? 400
13.2 Symbolic and Sacred Space 409
13.3 Myth and Religious Doctrine 412
13.4 Rituals of Transition and Conformity 416
13.5 Other Forms of Religious Practice 421
Key Terms 426
Summary 427
Critical Thinking Questions 428
Bibliography 428

CHAPTER 14
Anthropology of Food 431
Introduction 431
14.1 Food as a Material Artifact 432
14.2 A Biocultural Approach to Food 437
14.3 Food and Cultural Identity 442
14.4 The Globalization of Food 448
Key Terms 452
Summary 452
Critical Thinking Questions 453
Bibliography 453

CHAPTER 15
Anthropology of Media 457
Introduction 457
15.1 Putting the Mass into Media 458
15.2 Putting Culture into Media Studies 461
15.3 Visual Anthropology and Ethnographic Film 462
15.4 Photography, Representation, and Memory 465

Access for free at openstax.org


15.5 News Media, the Public Sphere, and Nationalism 468
15.6 Community, Development, and Broadcast Media 472
15.7 Broadcasting Modernity and National Identity 474
15.8 Digital Media, New Socialities 475
Key Terms 481
Summary 481
Critical Thinking Questions 481
Bibliography 482

CHAPTER 16
Art, Music, and Sport 485
Introduction 485
16.1 Anthropology of the Arts 486
16.2 Anthropology of Music 495
16.3 An Anthropological View of Sport throughout Time 502
16.4 Anthropology, Representation, and Performance 505
Key Terms 511
Summary 511
Critical Thinking Questions 511
Bibliography 512

CHAPTER 17
Medical Anthropology 515
Introduction 515
17.1 What Is Medical Anthropology? 516
17.2 Ethnomedicine 520
17.3 Theories and Methods 523
17.4 Applied Medical Anthropology 531
Key Terms 540
Summary 541
Critical Thinking Questions 541
Bibliography 542

CHAPTER 18
Human-Animal Relationship 545
Introduction 545
18.1 Humans and Animals 546
18.2 Animals and Subsistence 552
18.3 Symbolism and Meaning of Animals 556
18.4 Pet-Keeping 559
18.5 Animal Industries and the Animal Trade 562
Key Terms 569
Summary 569
Critical Thinking Questions 570
Bibliography 570

CHAPTER 19
Indigenous Anthropology 573
Introduction 573
19.1 Indigenous Peoples 574
19.2 Colonization and Anthropology 583
19.3 Indigenous Agency and Rights 589
19.4 Applied and Public Anthropology and Indigenous Peoples 601
Key Terms 604
Summary 605
Critical Thinking Questions 605
Bibliography 605

CHAPTER 20
Anthropology on the Ground 609
Introduction 609
20.1 Our Challenging World Today 610
20.2 Why Anthropology Matters 614
20.3 What Anthropologists Can Do 618
Key Terms 624
Summary 624
Critical Thinking Questions 624
Bibliography 625

Index 627

Access for free at openstax.org


Preface 1

PREFACE
About OpenStax textbook, errors sometimes occur, and new
perspectives highlight areas that require revision.
OpenStax is part of Rice University, which is a
Writing style guides and other contextual
501(c)(3) nonprofit charitable corporation. As an
frameworks also change frequently. Since our books
educational initiative, it’s our mission to transform
are web-based, we can make updates periodically
learning so that education works for every student.
when deemed pedagogically necessary. If you have a
Through our partnerships with philanthropic
correction to suggest, submit it through the link on
foundations and our alliance with other educational
your book page on OpenStax.org. Subject matter
resource companies, we’re breaking down the most
experts review all errata suggestions. OpenStax is
common barriers to learning. Because we believe
committed to remaining transparent about all
that everyone should and can have access to
updates, so you will also find a list of past errata
knowledge.
changes on your book page on OpenStax.org.
About OpenStax Resources Format
Customization You can access this textbook for free in web view or
Introduction to Anthropology is licensed under a PDF through OpenStax.org, and for a low cost in
Creative Commons Attribution 4.0 International (CC print.
BY) license, which means that you can distribute,
About Introduction to Anthropology
remix, and build upon the content, as long as you
provide attribution to OpenStax and its content Introduction to Anthropology is a four-field text,
contributors. grounded in foundational content in cultural
anthropology, archaeology, biological anthropology,
Because our books are openly licensed, you are free
and linguistic anthropology. This approach makes
to use the entire book or select only the sections that
the text useful for both general and cultural
are most relevant to the needs of your course. Feel
introductory courses as well as for introductory
free to remix the content by assigning your students
courses in some of the anthropology subfields. Upon
certain chapters and sections in your syllabus, in the
this strong foundation, two contemporary themes
order that you prefer. You can even provide a direct
are highlighted: social inequality and the natural
link in your syllabus to the sections in the web view
world. Ethnographies and examples throughout the
of your book.
text address the impacts of these two themes on
Instructors also have the option of creating a human societies throughout history and around the
customized version of their OpenStax book. The globe.
custom version can be made available to students in
Coverage and Scope
low-cost print or digital form through their campus
Introduction to Anthropology contains all of the
bookstore. Visit the Instructor Resources section of
foundational material necessary for introductory
your book page on OpenStax.org for more
courses in anthropology. Methods and theories from
information.
all four fields are introduced in the first two chapters
Art attribution and woven throughout later discussions. The central
In Introduction to Anthropology, art contains concept of culture likewise is both explored in detail
attribution to its title, creator or rights holder, host in its own chapter and referenced repeatedly in
platform, and license within the caption. Because examples throughout the text. The evolution and
the art is openly licensed, anyone may reuse the art diversification of the human species is centrally
as long as they provide the same attribution to its featured in two chapters, “Biological Evolution and
original source. Early Human Evidence” and “Physical and Cultural
Evolution in the Genus Homo.” The breadth of the
Errata
discipline is apparent in the variety of examples and
All OpenStax textbooks undergo a rigorous review ethnographies as well as specific chapters dedicated
process. However, like any professional-grade to developing areas of anthropology, such as
2 Preface

“Medical Anthropology” and “Human-Animal working, are likewise representative of the growing
Relationships.” An engaging and inviting narrative diversity of the field.
will hold students’ interest.
Unique chapters: Five of the text’s 20 chapters
Addressing Societal Issues introduce students to current and developing
The central themes of Introduction to specializations within the discipline. These chapters
Anthropology—social inequality and the natural offer an engaging and in-depth look at research
world—connect the text’s foundational material to fields rarely covered in introductory texts, fields that
two of the most pressing contemporary issues facing are particularly interdisciplinary in their aims and
societies around the world. practices. They further stress that anthropology is
an evolving and relevant field, offering insights into
• In addressing social inequality, the text drives
humanity’s deepest questions and directions
readers to consider the rise and impact of social
forward in addressing the toughest challenges.
inequalities based on forms of identity and
These chapters are:
difference (such as gender, ethnicity, race, and
class) as well as oppression and discrimination. • “Anthropology of Food,” including material on
The contributors to and dangers of food artifacts, ancient foodways and food
socioeconomic inequality are fully addressed, reconstructions, food as cultural heritage, food
and the role of inequality in social dysfunction, prescriptions and proscriptions, and the
disruption, and change is noted. Introduction to globalization of food.
Anthropology centers on the lived experiences • “Anthropology of Media,” addressing topics such
of a wide range of people and provides ample as visual anthropology and ethnographic film,
opportunities for instructors and students to photography and representation, news media
discuss and address preconceived notions, and the public sphere, the role of media in the
misconceptions, and potential solutions and development of national identity, and digital
outcomes. media.
• To illustrate the fundamental relationship • “Medical Anthropology,” with material on the
between humans and their environments, the history of medical anthropology, the social
natural world is treated as both a setting for construction of health, common medical
human existence and a key influence on human anthropology methods and theoretical
culture, economics, and politics. This focus approaches, and applied medical anthropology.
makes the text uniquely suited to the • “Human-Animal Relationships,” including
contemporary era as climate change and discussions of multispecies ethnography,
environmental degradation play an increasing human-animal empathy, human-animal
role in humanity’s governance, intercultural relationships among people practicing varying
relationships, and daily lives. subsistence strategies, animal symbolism in
oral tradition and religion, and pet keeping.
Illuminating an Evolving and Relevant Field
• “Indigenous Anthropology,” which, through the
The text showcases the historical context of the lens of the experiences of the Indigenous
discipline, with a strong focus on anthropology as a peoples of North America, addresses the
living and evolving field. A deep and reflective historical and contemporary challenges facing
exploration of the origins of anthropology’s methods Indigenous people, including issues of agency,
and goals is featured in several chapters, including rights, and identity, as well as exploring
“Methods: Cultural and Archaeological Research Indigenous material cultures, perspectives, and
Methods” and “Indigenous Anthropology.” There is worldviews.
significant discussion of recent efforts to make the
field more diverse—in its practitioners, in the Enriching and Engaging Features
questions it asks, and in the applications of Several feature boxes highlight the vibrant and
anthropological research to address contemporary applied nature of anthropology and give students
challenges. The authors who contributed to this text practice using the methods discussed throughout
come from diverse backgrounds and geographic the text.
regions, providing balance and richness to the
• Profiles in Anthropology. Each chapter
narrative, examples, and theoretical foundations of
contains a profile of one or more
the text. The researchers highlighted in the Profiles
anthropologists, many contemporary and some
in Anthropology sections, many still living and

Access for free at openstax.org


Preface 3

historical, who have made significant About the Authors


contributions to the discipline. These featured
Senior Contributing Authors
anthropologists represent a diversity of racial
Jennifer Hasty is an adjunct professor of African
and ethnic backgrounds as well as a broad
studies at the University of Pennsylvania. She
sampling of research interests and perspectives.
studies media and politics in West Africa and the
• Ethnographic Sketches. Ethnographic sketches
United States. Her book The Press and Political
taken from the authors’ own fieldwork are
spaced throughout the book. These engaging Culture in Ghana explores the cultural and historical
forces shaping the practice of journalism in the
vignettes provide a window into the actual work
recent period of democratization. In addition to
of doing anthropology, providing readers with a
working as a journalist for several Ghanaian media
sense of the pleasures and challenges of doing
organizations, she has worked as a wedding
research in the field.
videographer in the Philadelphia metro area and a
• Mini-Fieldwork/Applied Activities. Each
community radio DJ in northern New Mexico. She is
chapter concludes with a simple fieldwork
currently writing a book on corruption in Ghana.
activity to give students practice thinking and
Chapters authored or coauthored in this text include
researching like an anthropologist. These
the following:
exercises provide them with hands-on
experience applying the methods and theories Chapter 1: What Is Anthropology?
discussed in the chapter to actual research
Chapter 2: Methods: Cultural and Archaeological
conducted in their own communities.
Chapter 3: Culture Concept Theory: Theories of
Pedagogical Framework
Cultural Change
An effective pedagogical framework helps students
structure their learning and retain information. Chapter 6: Anthropological Thought

• Chapter Outlines. Each chapter opens with an Chapter 7: Work, Life, and Value: Economic
outline and introduction, familiarizing students Anthropology
with the material that will follow. Throughout
Chapter 8: Authority, Decisions, and Power: Political
the chapter, material is chunked into
Anthropology
manageable sections of content within each of
the larger main heads. Chapter 12: Gender and Sexuality
• Learning Objectives. Every main section begins
David G. Lewis is a member of the Confederated
with a set of clear and concise learning
Tribes of Grand Ronde of Oregon. He has a PhD from
objectives. These objectives are designed to help
the University of Oregon (2009) and is an assistant
the instructor decide what content to include or
professor of anthropology and ethnic studies at
assign and to guide student expectations. After
Oregon State University. David has conducted
completing the section and relevant end-of-
research on Oregon tribal history for some 25 years
chapter exercises, students should be able to
and has published numerous journal articles and
demonstrate mastery of the learning objectives.
book chapters. Additionally, he has researched and
• Chapter Summaries. Chapter summaries distill
written over 470 essays for his blog, the Quartux
the information presented in each chapter to
Journal (https://openstax.org/r/QuartuxJournal),
key, concise points.
documenting tribal adjustments to colonization in
• Key Terms. Key terms are bolded and followed
the West. David conducts numerous presentations
by a definition within the text. Definitions of key
annually with community groups, at conferences,
terms are also listed in a glossary at the end of
and at universities, educating about tribes in the
each chapter.
region; consults with local governments and
• Critical Thinking Questions. Each chapter
organizations on diversity, place naming, and land
ends with 8 to 10 critical thinking questions
acknowledgments; and curates museum exhibits at
designed to help students assess their learning
local historical societies and museums. Chapters
and apply it to their daily lives.
authored or coauthored in this text include the
• Suggested Readings. This feature helps
following:
students further explore the chapter content by
providing curated links to other information Chapter 2: Methods: Cultural and Archaeological
sources.
Chapter 3: Culture Concept Theory: Theories of
4 Preface

Cultural Change Tony Fitzpatrick, University of Wyoming

Chapter 19: Indigenous Anthropology Paul Hanson, Case Western Reserve University

Dr. Marjorie M. Snipes earned a PhD in cultural David Hicks, Stony Brook University
anthropology from the University of
Michael Hollis, St. Edward’s University
Wisconsin–Madison (1996) and is currently a
professor of anthropology at the University of West Stewart Jobrack, The Ohio State University at
Georgia, where she teaches anthropological theory, Mansfield
ethnographic field methods, anthropology of
Barry Kass, Orange County Community College
religion, and animals and culture. Her doctoral
fieldwork in the northwestern Andes of Argentina Phineas Kelly, University of Wyoming
focused on religion and identity in an agropastoral
Elizabeth Kickham, Idaho State University
society, in particular on understanding the
relationships that herders forge with their animals Jonathan Marion, University of Arkansas
and with each other. Among her recent publications
Annie Melzer, Northern Kentucky University
are Inside Anthropology (2021, Kendall Hunt) and
The Intellectual Legacy of Victor and Edith Turner Kerith Miller, University of Arizona
(2018, Lexington). Chapters authored or coauthored
Mackie O’Hara, The Ohio State University
in this text include the following:
Jenell Paris, Messiah University
Chapter 10: The Global Impact of Human Migration
Caroline Rivera, Florida Gulf Coast University
Chapter 11: Forming Family through Kinship
Megan Schmidt-Sane, Case Western Reserve
Chapter 13: Religion and Culture
University
Chapter 14: Anthropology of Food
Max Stein, Florida Gulf Coast University
Chapter 18: Human-Animal Relationships
Fay Stevens, University of Notre Dame
Chapter 20: Anthropology on the Ground
Antoaneta Tileva, American University
Contributing Authors
Kristen Verostick, Rowan University
Dr. Todd A. Barnhardt, PhD
Additional Resources
Dr. M. Anne Basham, Gateway Community College
and Executive Director, Biodiversity Outreach Student and Instructor Resources
Network We’ve compiled additional resources for both
students and instructors, including an instructor’s
Sharon Gursky, Texas A&M University
manual, test bank, and lecture slides. Instructor
Laura Jarvis-Seibert resources require a verified instructor account,
which you can apply for when you log in or create
Saira A. Mehmood
your account on OpenStax.org. Take advantage of
Dr. Sydney Yeager, Rollins College these resources to supplement your OpenStax book.

Reviewers Comprehensive Instructor’s Manual. Each


Janet Altamirano, Texas A&M University–Kingsville component of the instructor’s manual is designed to
provide maximum guidance for delivering the
Dr. M. Anne Basham, Gateway Community College
content in an interesting and dynamic manner. The
Jack Bish, University of Wisconsin–Madison instructor’s manual includes a chapter outline
containing the learning outcomes for each section,
Heidi Bludau, Monmouth University
section outlines, and section summaries. Chapter
Ryan Collins, Dartmouth College key terms are listed as well. Also included for each
chapter are strategies for using the Mini-Fieldwork/
Alejandra Dashe, Paradise Valley Community College
Applied Activity and the Profiles in Anthropology.
Bridget Fitzpatrick, Normandale Community College There are sample answers and strategies for using
select critical thinking questions in the chapter.
Leslie Fitzpatrick, Mercyhurst University
Each chapter also includes links to websites and

Access for free at openstax.org


Preface 5

organizations relevant to the content in the chapter resources that support OpenStax books, free of
as well as to content that extends examples in the charge. Through our Community Hubs, instructors
chapter. can upload their own materials or download
resources to use in their own courses, including
Test Bank. With nearly 1,100 multiple-choice, fill-
additional ancillaries, teaching material,
in-the-blank, and short-answer questions in the test
multimedia, and relevant course content. We
bank, instructors can customize tests to support a
encourage instructors to join the hubs for the
variety of course objectives. The test bank is
subjects most relevant to your teaching and
available in Word format.
research as an opportunity both to enrich your
PowerPoint Lecture Slides. The comprehensive courses and to engage with other faculty. To reach
PowerPoint lecture slides provide a structure for the Community Hubs, visit www.oercommons.org/
course lectures. Chapter images, lesson learning hubs/openstax.
outcomes, and bulleted content provide a starting
Technology Partners
place for instructors to build their lectures.
As allies in making high-quality learning materials
Community Hubs accessible, our technology partners offer optional
OpenStax partners with the Institute for the Study of low-cost tools that are integrated with OpenStax
Knowledge Management in Education (ISKME) to books. To access the technology options for your
offer Community Hubs on OER Commons—a text, visit your book page on OpenStax.org.
platform for instructors to share community-created
6 Preface

Access for free at openstax.org


CHAPTER 1
What Is Anthropology?

Figure 1.1 Artist’s depiction of a woman hunting, created in 1565. Contrary to some long held beliefs, women have
always played a role in hunting game. (credit: “Illustration of activities of Lapps and Finns: Men and women hunting
with bows and arrows on snowshoes; “women hunt...as nimbly...or more than men” by Illustration of activities of
Lapps and Finns/Library of Congress Prints and Photographs Division)

CHAPTER OUTLINE
1.1 The Study of Humanity, or "Anthropology Is Vast"
1.2 The Four-Field Approach: Four Approaches within the Guiding Narrative
1.3 Overcoming Ethnocentrism
1.4 Western Bias in Our Assumptions about Humanity
1.5 Holism, Anthropology’s Distinctive Approach
1.6 Cross-Cultural Comparison and Cultural Relativism
1.7 Reaching for an Insider’s Point of View

INTRODUCTION Imagine a research project that contains these three members:

Randy Haas discovered the 9,000-year-old grave of a teenager buried with a hunting tool kit in the Andes
mountains of Peru. Haas found that this hunter from long ago was a young woman. This discovery has upset
the notion that hunting was the exclusive activity of men throughout human evolutionary history.

Daniel Miller is part of a global team researching how people use smartphones in various parts of the world,
including Brazil, Cameroon, Chile, China, Ireland, Italy, Japan, East Jerusalem, and Uganda. The team is
exploring how smartphones take on different functions in different cultural contexts. Focusing on Ireland,
8 1 • What Is Anthropology?

Miller theorizes that smartphones become a kind of personal avatar, expressing and enacting the specific
social identity of the user.

FIGURE 1.2 Red-tailed monkeys, the subject of anthropologist Michelle Brown’s study, are primates that are found
in Central and East Africa. This red-tailed monkey lives in Uganda. They are social animals and live in groups of 8-30
individuals. (credit: “Schmidt's Red-tailed Monkey” by Mehgan Murphy/Smithsonian’s National Zoo, CC0 1.0)

Michelle Brown spends long days observing blue monkeys, red-tailed monkeys, and baboons in a conservation
park in Uganda. She records the behavior of these primates as they find food, communicate, and fight with one
another. She collects urine and feces to analyze hormone levels, intestinal parasites, and DNA. She wants to
understand how primates compete as individuals and groups for access to various foods in their environment.

What kind of research project could encompass such a diversity of topics and methods? Since this is the first
chapter of an anthropology textbook, you can probably guess. Though they conduct research on vastly
different topics, all three are anthropologists. How could the work of these researchers be united in one
academic discipline? The reason, as we will see, is that anthropology is vast.

Anthropology, the study of humanity, is guided by a central narrative and set of research commitments.
Anthropology aims to overcome bias by examining cultures as complex, integrated products of specific
environmental and historical conditions. Anthropologists use many different research strategies in their
efforts to represent people from cultures very different from their own.

Anthropology explores controversial topics that may challenge individual assumptions and values. The goal is
to understand the full experience of humanity, including elements that may seem unfamiliar or
uncomfortable. Anthropology teaches a set of skills for setting aside personal perspectives and keeping an
open mind while learning about the diversity of human practices and ideas. As discussed further at the end of
this chapter, this does not mean abandoning individual personal values, but rather suspending judgment
temporarily while learning to understand the perspectives of others.

1.1 The Study of Humanity, or "Anthropology Is Vast"


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define the study of anthropology in the broadest sense.
• Summarize the guiding narrative of anthropology.
• Restate and explain the central commitments of anthropology.

Anthropology is a vast field of study—so vast, in fact, that anthropology is interested in everything.
Anthropology is unique in its enormous breadth and its distinctive focus. Consider other disciplines. In the
arts and sciences, each discipline focuses on a discrete field of social life or physical phenomena. Economists
study economics. Religious scholars study religion. Environmental scientists study the environment.

Access for free at openstax.org


1.1 • The Study of Humanity, or "Anthropology Is Vast" 9

Biologists study living organisms. And so on.

Anthropologists study all of these things. Put simply, anthropology is the study of humanity across time and
space. Anthropologists study every possible realm of human experience, thought, activity, and organization.
Human as we are, we can only engage in social and natural worlds through our human minds and human
bodies. Even engagement with nonhuman realms such as astronomy and botany is conditioned by our human
senses and human cognition and thus varies across different societies and different time periods.

You may be thinking, If anthropology is the human aspect of absolutely everything, then does anthropology
encompass the other social disciplines, such as political science, religious studies, and economics? This is not
the case. Certainly, anthropologists are frequently multidisciplinary, meaning that while their research and
teaching are focused within the discipline of anthropology, they also engage with other disciplines and work
with researchers and teachers in other fields. But the way that scholars in the other social disciplines approach
their subject matter is different from the way anthropologists approach those same subjects.

The distinctive approach of anthropology relies on a central narrative, or story, about humanity as well as a set
of scholarly commitments. This central story and these common commitments hold the discipline together,
enabling anthropologists to combine insights from diverse fields into one complex portrait of what it means to
be human.

Anthropology is everything, but it’s not just anything. Anthropology is the study of humanity guided by a
distinctive narrative and set of commitments.

The Heart of Anthropology: Central Narrative and Commitments


Anthropologists are great storytellers. They tell many, many stories about all aspects of human life. At the
heart of all of these stories is one fundamental story: the “story of humanity,” a rich and complex narrative. A
narrative is a story that describes a connected set of features and events. Narratives can be fictional or
nonfictional. The narrative of anthropology is a true story, a factual narrative about the origins and
development of humanity as well as our contemporary ways of life. The central narrative of anthropology can
be summarized this way.

Human beings have developed flexible biological and social features that have worked together in a wide
variety of environmental and historical conditions to produce a diversity of cultures.

Three features of this narrative are especially important to anthropologists. These features form three central
commitments of anthropology. In academic study, a commitment is a common goal recognized by the scholars
in a discipline.

Central Commitment #1: Exploring Sociocultural Diversity


As the narrative suggests, humans in a diversity of conditions create a diversity of cultures. Rather than trying
to find out which way of life is better, morally superior, more efficient, or happier or to make any other sort of
judgment call, anthropologists are committed to describing and understanding the diversity of human ways of
life. Setting aside judgments, we can see that humans everywhere create culture to meet their needs.
Anthropologists discover how different cultures devise different solutions to the challenges of human survival,
social integration, and the search for meaning.

What are you wearing today? Perhaps a T-shirt and jeans with sneakers, or a tunic and leggings with flip-flops.
What about your professor? Are they wearing a bathrobe and slippers, or perhaps a cocktail dress with
stilettos? You can be (almost) certain that will never happen. But why not? You might assume that what
Americans wear for class is completely normal, but this assumption ignores the question of what makes
something “normal.”
10 1 • What Is Anthropology?

FIGURE 1.3 Ghanaian professionals wearing local fashions to promote the national textile industry. (credit:
“Ghanaian Ladies” by Erik (HASH) Hersman/flickr, CC BY 2.0)

In many countries, for instance, university students typically wear dress shirts with slacks or skirts to class.
Many Ghanaian students would not dream of wearing ripped jeans or tight leggings to class, considering such
casual dress disrespectful. American students put much more emphasis on comfort than on presentation, an
overall trend in American dress. Even in office settings, it is now acceptable for Americans to wear casual
clothing on Fridays. In the West African country of Ghana, “casual Friday” never caught on, but office workers
have developed their own distinctive Friday dress code. As the local textile industry became threatened by
Chinese imports, Ghanaian office workers began wearing outfits sewn from locally manufactured cloth on
Fridays, creating a practice of “National Friday Wear.”

So which way is better, the American way or the Ghanaian way? Anthropologists understand that neither way is
better and that each addresses a need within a particular culture. Casual Friday is great for Americans who
crave comfy leisurewear, while National Friday Wear is great for Ghanaians who want to boost their local
economy and show their cultural pride.

Anthropologists recognize not only diversity across different cultures but also the diverse experiences and
perspectives within a culture. Do you ever buy used clothing at thrift shops, or do you know people who do? An
old green men’s trench coat bought at a vintage clothing store may be a favorite of a college student. The
mother of that student may not feel the same way and offer to buy their child a new coat much to the distress of
the owner of the coat! To people who have grown up in the 1930’s and 1940’s, used clothing was associated
with the hard times of the Great Depression. For the newer generations, used clothing is a way to find unique,
affordable clothing that can stretch the boundaries of mainstream style. Although people in a culture share a
general set of rules, they interpret them differently according to their social roles and experiences, sometimes
stretching the rules in ways that ultimately change them over time.

Access for free at openstax.org


1.1 • The Study of Humanity, or "Anthropology Is Vast" 11

FIGURE 1.4 Ghanaian trader in her secondhand clothing shop (credit: “Market Woman - Kejetia Market - Kumasi -
Ghana - 03” by Adam Jones/flickr, CC BY 2.0)

In Ghana, most used clothing is imported from the United States and Europe in large bales that local vendors
purchase and sell in market stalls. A person from the United States or Europe is locally referred to as an
obruni. Used clothing is called obruni wawu, or “a foreign person has died,” reflecting the assumption that no
living person would give away such wearable clothing. Many Ghanaians love to pick through the piles of obruni
wawu in the market, thrilled to find recognizable brands and unusual styles. Some, however, associate obruni
wawu with poverty. The stalls that sell obruni wawu are often called “bend-over boutiques,” referring to the
subservient posture adopted by customers rifling through the piles of clothing on the ground. Obruni wawu is
suitable in some situations but certainly not in others. A particular Ghanaian movie included a scene where a
man trying to woo a much younger woman. When the man gave his would be girlfriend a bag full of obruni
wawu as a gift, it caused the audience to burst our laughing. The gift was humorous and inappropriate to the
audience.

As with clothing, different cultures come up with different solutions to common challenges such as housing,
food, family structure, the organization of work, and finding meaning in life. And people in every society
discuss and argue about their own cultural norms. Anthropology seeks to document and understand the
diverse range of solutions to common human challenges as well as the diversity of conflicting perspectives
within each culture.

Central Commitment #2: Understanding How Societies Hold Together


Just as the various parts of our bodies all work together (the brain, the heart, the liver, the skeleton, and so
forth), the various parts of a society all work together as well (the economy, the political system, religion,
families, etc.). Frequently, anthropologists discover that changes in one realm of society are related to changes
in another realm in unexpected ways. When farmers in Ghana began growing cocoa for export during the
colonial period, the agricultural shift dramatically altered gender relations as men monopolized cash crops
and women were relegated to vegetable farming for their families’ consumption and local trade. As men
benefited from the profits of the cocoa trade, relations between men and women became more unequal.

Anthropologists have a favorite word for the way that all elements of human life interrelate to form distinctive
12 1 • What Is Anthropology?

cultures: holism. Sometimes those parts reinforce one another, encouraging stability; sometimes they
contradict one another, promoting change. Consider the caste system in India. Cultural anthropologist Susan
Bayly describes how the beliefs and practices associated with caste in India have provided cultural integration
and stability while also demonstrating a great deal of local variability and working as a force of social change
(1999). Most Indians are familiar with two forms of belonging assigned by birth, the jati (birth group) and the
varna (order, class, or kind). There are thousands of birth groups in the various regions of India, many specific
to a single region. By contrast, there are four varnas known across India: Brahmins (associated with priests),
Kshatriyas (associated with rulers and warriors), Vaishyas (associated with traders), and Shudras (associated
with servile laborers). Another group, called “untouchables” or dalits, are outside the scheme of varnas.

As described in the Vedas, the four varnas are ordered in an interdependent hierarchy reminiscent of human
anatomy. The Rig Veda describes how the gods sacrificed the first man, Purusa, dividing his body to create four
groups of humanity:

When they divided the Purusa, into how many parts did they arrange him? What was his mouth? What
his two arms? What are his thighs [loins] and feet called? The brahmin was his mouth, his two arms
were made the rajanya [kshatriya, king and warrior], his two thighs [loins] the vaisya, from his feet the
sudra [servile class] was born. (Bayle, 1999)

Ancient texts envision caste as a means of social order as people in each caste perform different functions and
occupations, all working together in harmony. Note, however, that such texts were written down by members of
upper-caste groups, often Brahmin scholars. Anthropologists and historians who study the practices of caste
argue that the caste system was never such a unitary and dominant force across the country but rather a
flexible, regional, and constantly changing set of identities. In the colonial period, the British made the caste
system more rigid and antagonistic, offering education and jobs to select caste groups. In the 20th century,
many lower-caste groups have resisted their oppression by converting to Christianity or Islam and forming
political parties to pressure the government for more opportunities for social advancement.

Anthropologists are curious about how different cultures create different categories of people and use those
categories to organize the activities of social life. In many farming societies, for instance, men do certain kinds
of agricultural work and women do others. In societies where land must be cleared in order to sow crops, men
often chop down trees and clear the brush while women do the planting. In societies that utilize large-scale
industrial farming, migrants or people of a specific ethnicity or assigned racial category are often recruited (or
forced) to perform the manual labor required to grow and harvest crops. In industrial capitalist societies, one
group of people owns the factories and another group works the machines that produce the industrial
products. Relations between groups can be cooperative, competitive, or combative. Some cultures promote the
equality of social groups, while many others reinforce inequality among groups. Holism is not the same as
harmony. Anthropologists are interested in how society holds together but also in the conditions that can
cause conflict, change, and disintegration.

You may have heard the word polarized used to describe the sense that two different groups in American
society are moving farther and farther apart in their values, opinions, and desires. Some suggest that the
contradictory perspectives of these two groups threaten to tear American society apart. Others suggest that
Americans are united by deeper values such as freedom, equal opportunity, and democracy. Using holism to
understand this issue, an anthropologist might consider how the perspectives of each group relate to that
group’s economic experiences, political convictions, and/or religious or moral values. A comprehensive use of
holism would explore all of these aspects of society, looking at how they interact to produce the polarization we
see today and suggesting what might be done to bring the two groups into productive dialogue.

Central Commitment #3: Examining the Interdependence of Humans and Nature


As our narrative suggests, anthropologists are interested in the natural environment, the way humans have
related to the natural world over time, and how this relationship shapes various cultures. Anthropologists
consider how people in different cultures understand and use the various elements of nature, including land,
water, plants, animals, climate, and space. They show how people interact with these elements of nature in
complex ways.

Access for free at openstax.org


1.1 • The Study of Humanity, or "Anthropology Is Vast" 13

Archaeologists working in prehistoric sites all over the world have documented how prehistoric people
understood celestial objects and used them to navigate their waterways, create calendars and clocks, regulate
farming activities, schedule religious ceremonies, and inform political leaders. This area of study is called
archaeoastronomy. In Chaco Canyon in the American Southwest, archaeologists have discovered that
buildings in the major settlement areas were aligned so that certain windows would provide perfect vantage
points to view the sun and moon at pivotal times of the year, such as solstice and equinox. The Sun Dagger,
consisting of two whorl-shaped petroglyphs (stone etchings) on Fajada Butte, is precisely positioned under a
rock crevice so as to indicate the solstices and equinoxes when the sun shines through the crevice.
Unfortunately, tourist foot traffic at the site has altered the width and direction of the crevice so that the Sun
Dagger no longer marks these celestial events accurately.

FIGURE 1.5 The Sun Dagger at Fajada Butte (credit: National Park Service/Wikimedia Commons, Public Domain)

The people of Chaco Canyon may have been particularly attuned to the features of their environment as they
constructed their complex civilization in the challenging environment of the high desert. With scarce rainfall
and brief growing seasons, their survival depended on accurate identification of opportune planting and
harvesting times. With the onset of a 50-year drought, farming became more and more precarious. Eventually,
the ancient peoples of Chaco were forced to abandon the area.

Some anthropologists study how people interact with the plants in their area. The field of ethnobotany
examines how people in different cultures categorize and use plants for food, shelter, tools, transportation, art,
and religion. Ethnobotanists also conduct research on plants used in healing to discover the relationship
between cultural practices and the pharmaceutical properties of these plants. Some examine the cultural use
of psychoactive plants such as mushrooms and peyote in religious ritual. For instance, anthropologist Jamon
Halvaksz studied the controversial use of marijuana among youth in New Guinea (2006). Young people told
Halvaksz that marijuana helped them work harder, overcome shame, and understand ancestral stories. Critics
of the practice told Halvaksz that marijuana dried the blood of people who used it, making their offspring weak
and feeble. Marijuana use has generated similar controversies in other countries, including the United States,
with some arguing that the drug provides relaxation and pain relief while others claim it interferes with
cognitive abilities and motivation.

Our relationship with nature is reciprocal. Nature shapes humanity, and humanity shapes nature. Exploring
14 1 • What Is Anthropology?

how nature shapes humanity, anthropologists speculate about how aspects of the environment have shaped
the emergence and development of human biology, such as our ability to walk, the shape of our teeth, and the
size of our brains. Dramatic climactic shifts over the past several million years have forced periods of rapid
biological and cultural adaptation, resulting in new hominin species and new skill sets such as language and
toolmaking. In more recent archaeological time periods, environmental characteristics have shaped religious
beliefs, gender relations, food-getting strategies, and political systems. Environmental forces can trigger the
beginning or the end of a society. Some archaeologists study how natural events such as volcanic eruptions
and droughts have led to mass migrations and the collapse of empires.

Our reciprocal relationship with nature also works the other way around; that is, humans shape nature. Our
environments are shaped by the food-getting methods of our societies as well as the way we acquire and trade
resources such as oil, natural gas, diamonds, and gold. Many anthropologists explore how contemporary ways
of life change the natural world at local, regional, and global levels. Farming dramatically impacts ecosystems
with the clearing of prairies, wetlands, and forests. Fishing can deplete certain species, changing the whole
ecosystem of rivers and coastal waters. Responding to population pressures, people construct dams to channel
water to emergent cities. The redirection of water transforms regional ecosystems, turning wetlands into
deserts and deserts into resource-hungry cities.

Scholars use the term Anthropocene to describe the contemporary period of increasing human impact on the
ecosystems of our planet. Large-scale pollution, mining, deforestation, ranching, and agriculture are causing
dramatic environmental disruptions such as climate change and mass extinction of plant and animal species.
Many anthropologists are studying these problems, focusing on how people are working locally, regionally, and
globally to promote more sustainable ways of living in our natural world.

1.2 The Four-Field Approach: Four Approaches within the Guiding Narrative
LEARNING OUTCOMES

By the end of this section, you will be able to:


• Identify and define the four fields of anthropology.
• Describe the work of professional anthropologists in each field.
• Provide an example of how the four fields work together to explore common issues.

Let’s recall the central narrative of anthropology:

Human beings have developed flexible biological and social features that have worked together in a wide
variety of environmental and historical conditions to produce a diversity of cultures.

Researching this argument is a vast endeavor requiring many complementary approaches and techniques.
Anthropology comprises four main approaches, the four subfields of our discipline. Each subfield specializes
in exploring a different aspect of the common narrative. Combining insights from the four fields gives us a rich
and complex understanding of specific issues such as gender, inequality, race, and the environment. Let’s take
a look at each subfield and then examine how the subfields combine in the study of racial categories and
relations.

Biological Anthropology
Biological anthropology focuses on the earliest processes in the biological and sociocultural development of
human beings as well as the biological diversity of contemporary humans. In other words, biological
anthropologists study the origins, evolution, and diversity of our species. Some biological anthropologists use
genetic data to explore the global distribution of human traits such as blood type or the ability to digest dairy
products. Some study fossils to learn how humans have evolved and migrated. Some study our closest animal
relatives, the primates, in order to understand what biological and social traits humans share with primates
and explore what makes humans unique in the animal world.

The Dutch primatologist Carel van Schaik spent six years observing orangutans in Sumatra, discovering that
these reclusive animals are actually much more social than previously thought (2004). Moreover, van Schaik
observed that orangutans use a wide variety of tools and pass down skills to their young. By studying these

Access for free at openstax.org


1.2 • The Four-Field Approach: Four Approaches within the Guiding Narrative 15

primates, van Schaik and other biological anthropologists gain insight into the origins of human intelligence,
technology, and culture. These researchers also warn that habitat loss, illegal hunting, and the exotic pet trade
threaten the survival of our fascinating primate cousins.

Biological anthropologists frequently combine research among primates with evidence from the human fossil
record, genetics, neuroscience, and geography to answer questions about human evolution. Sometimes their
insights are startling and unexpected. Anthropologist Lynne Isbell argues that snakes have played a key role in
the evolution of human biology, particularly our keen sense of sight and our ability to communicate through
language (Isabell, 2009). Isbell’s “snake detection theory” posits that primates developed specialized visual
perception as well as the ability to communicate what they were seeing in order to alert others to the threat of
venomous snakes in their environment. She points to the near-universal fear of snakes shared by both
humans and primates and has documented the prevalence of snake phobia in human myth and folklore.
Isbell’s research highlights how human-animal relations are central to humanity, shaping both biology and
culture.

Not all biological anthropologists study primates. Many biological anthropologists study fossilized remains in
order to chart the evolution of early hominins, the evolutionary ancestors of modern humans. In this field of
study, anthropologists consider the emergence and migration of the various species in the hominin family tree
as well as the conditions that promoted certain biological and cultural traits. Some biological anthropologists
examine the genetic makeup of contemporary humans in order to learn how certain genes and traits are
distributed in human populations across different environments. Others examine human genetics looking for
clues about the relationships between early modern humans and other hominins, such as Neanderthals.

Forensic anthropology uses the techniques of biological anthropology to solve crimes. By analyzing human
remains such as decomposed bodies or skeletons, or tissue samples such as skin or hair, forensic
anthropologists discern what they can about the nature of a crime and the people involved. Key questions are
who died, how they died, and how long ago they died. Often, forensic anthropologists can discover the age, sex,
and other distinctive features of perpetrators and victims. Looking closely at forms of bodily trauma and
patterns of blood or bullets, they piece together the story of the crime. They work on investigative teams with
law enforcement officers and medical experts in ballistics, toxicology, and other specialties. Forensic
anthropologists often present their findings as witnesses in murder trials.

Not all of these crimes are contemporary. Sometimes, forensic anthropology is used to understand historical
events. Excavating the historic Jamestown colony of early English settlers in North America, archaeologist
William Kelso found a human skull in the midst of food remains. Noticing strange cut marks on the skull, he
called upon Douglas Owsley, a forensic anthropologist working for the Smithsonian Institution, to help him
figure out what the markings meant. Owsley determined that the markings were evidence of intentional
chopping to the skull with a sharp blade. He concluded that the skeleton belonged to a 14-year-old girl who had
been cannibalized by other settlers after she died. This interpretation corroborates historical evidence of
severe starvation in the colony during the harsh winter of 1609–1610.

Archaeology
Archaeologists use artifacts and fossils to explore how environmental and historical conditions have produced
a diversity of human cultures – the study of archaeology. Artifacts are objects made by human beings, such as
tools or pottery. Fossils are the remains of organisms preserved in the environment. Archaeologists have
developed careful methods of excavation, or removing fossils and artifacts from the ground, in order to learn
as much as possible about how people lived in times before and after the development of writing. They are
interested in how people met basic needs such as clothing and shelter, as well how they organized their
societies in family groups, trade networks, and systems of leadership. Many archaeologists seek to understand
how humans lived in relation to the natural world around them, altering the environment at the same time that
the environment was shaping their evolution and social development.

A group of archaeologists led by Tom Dillehay spent seven years excavating a set of sites in northern Peru,
charting the development of human society in this area over a period of 14,000 years (2017). They traced the
society from the early ways of life to the emergence of cities and early states, discovering how people there
developed fishing, farming, and herding strategies that led to increased sociocultural complexity. The team
16 1 • What Is Anthropology?

collected data on the plants and animals of the area as well as the buildings, tools, cloth, and baskets made by
the people. They concluded that the people who lived in this area placed a high value on cooperation and living
in harmony with nature.

Some archaeologists focus on more specific topics in more recent time periods. Archaeologist Eric Tourigny
examined the graves at pet cemeteries in the United Kingdom from 1881 to 1981(2020). Looking at the
epitaphs on the gravestones of the pets, Tourigny noted a change from earlier Victorian ways of thinking of
pets as friends to later, more modern ways of conceptualizing pets as members of the family. He noted, too,
that epitaphs expressed an increasingly common belief that pet owners would be reunited with their pets in
the afterlife.

Cultural Anthropology
Cultural anthropology is devoted to describing and understanding the wide variety of cultures referred to in
anthropology’s central narrative. Cultural anthropologists explore the everyday thoughts, feelings, and actions
of people in different cultures as well as the cultural and historical events that they consider important.
Examining social discourse and action, cultural anthropologists seek to understand unspoken norms and
values as well as larger forces such as economic change and political domination. Cultural anthropologists also
study how different societies are structured, including the roles and institutions that organize social life.

Cultural anthropologists often live for many months or years in the societies they study, adopting local ways of
living, eating, dressing, and speaking as accurately as possible. This practice is called fieldwork.
Anthropologists who undertake fieldwork might write an ethnography, an in-depth study of the culture they
have been studying. Classic ethnographies of the early 20th century often portrayed the cultures of non-
Western peoples as harmonious and unchanging over time. Bronislaw Malinowski, a pioneer of the long-term
fieldwork method, spent nearly two years studying trade and magic among the Trobriand peoples living in
what is now the Kiriwina island chain northeast of New Guinea. His ethnography, Argonauts of the Western
Pacific (1922), describes how Trobrianders undertook canoe voyages from island to island for the ceremonial
exchange of white shell bracelets and red shell necklaces among different island groups, an exchange system
known as the kula ring. Curiously, these highly valued objects had no use whatsoever, as no one ever wore
them. Rather, the exchange of bracelets and necklaces functioned as a means of enhancing social status (for
the givers) and reinforcing trade relationships. Malinowski argues that this form of exchange took the place of
warfare. Exploring the kula ring in great detail, Malinowski also learned about many other aspects of
Trobriand culture, such as the making of tools and canoes, farming practices, gender roles, sexuality, and
magical beliefs and practices.

Nowadays, cultural anthropologists tend to focus more on issues involving conflict and change, such as suicide
bombing in Afghanistan (Edwards 2017), a creationist theme park in Kentucky (Bielo 2018), sperm donation in
Denmark (Mohr 2018), and garbage pickers in Rio de Janeiro (Millar 2018). Often, anthropologists explore
overlooked and marginalized perspectives on controversial issues, shedding light on the cultural complexities
and power dynamics involved. Anthropologist Tracey Heatherington was interested in why some people were
resisting the creation of a conservation park on the Italian island of Sardinia (2010). The central highlands of
Sardinia are home to many endangered species and old growth forests, as well as local herding peoples who
fiercely resisted the appropriation of their homeland. Heatherington’s research identified three competing
perspectives: those of global environmentalists, the national government of Italy, and the local people of
Sardinia. The global environmentalists view the Sardinian highlands as a delicate ecosystem that should be
protected and controlled by environmental experts. The Italian government sees in the same land an
opportunity to develop ecotourism and demonstrate the Italian commitment to environmentalism. The local
peoples of Sardinia treasure their homeland as the foundation of their way of life, an intimate landscape
imbued with history and cultural value. As the controversy drew these three perspectives together, Western-
led global environmentalism combined with national government to undermine the legitimacy of local
knowledge and authority. Heatherington describes how stereotypes of Sardinians as ignorant and culturally
backward were used to delegitimize their resistance to the conservation park, drawing our attention to forms
of ecological racism that lurk in the global environmental movement.

Access for free at openstax.org


1.2 • The Four-Field Approach: Four Approaches within the Guiding Narrative 17

Linguistic Anthropology
As you might guess, linguistic anthropology focuses on language. Linguistic anthropologists view language as
a primary means by which humans create their diverse cultures. Language combines biological and social
elements. Some linguistic anthropologists study the origins of language, asking how language emerged in our
biological evolution and sociocultural development and what aspects of language might have given early
hominins an evolutionary advantage. Other linguistic anthropologists are interested in how language shapes
our thinking processes and our views of the world. In addition to its cognitive aspects, language is a powerful
tool for getting things done. Linguistic anthropologists also study how people use language to form
communities and identities, assert power, and resist authority.

Linguistic anthropologists frequently conduct the same kinds of long-term, immersive research that cultural
anthropologists do. Christopher Ball spent a year living and traveling with the Wauja, an indigenous group in
Brazil (2018). He describes the many routine and ritualized ways of speaking in this community and how each
kind of talk generates specific types of social action. “Chief speech” is used by leaders, while “bringing the
spirits” is used for healing the sick. Ceremonial language is used for giving people names and for conducting
exchanges between different indigenous groups. Ball, like many linguistic anthropologists, also examined
public speeches, such as the ones delivered by Wauja leaders to protest a dam on a nearby river. Ball also
analyzed the forms of language used by state officials and development workers to marginalize and
subordinate indigenous groups such as the Wauja.

Language is central to the way we conceptualize ourselves and our lives. Have you ever been asked to write an
essay about yourself, perhaps as part of a school assignment or college application? If so, you might have used
different phrases and concepts than if you’d been chatting with a new acquaintance. The purpose and intended
audience of our language use shapes the way we represent ourselves and our actions.

Anthropologist Summerson Carr examined an addiction treatment program for homeless women in the
midwestern United States, looking at the role of language in the therapeutic process (2011). After observing
therapy sessions and self-help meetings, she describes how addiction counselors promote a certain kind of
“healthy talk” that conveys deep cultural notions about personhood and responsibility. As patients master this
“healthy talk,” they learn to demonstrate progress by performing very scripted ways of speaking about
themselves and their addiction.

How the Four Fields Work Together: The Example of Race


With their unique methods and emphases, the four fields of anthropology may seem like completely different
disciplines. It’s true that anthropologists from the four fields don’t always agree on the best approach to
sociocultural enquiry. Biological anthropologists often see themselves as “hard” scientists committed to
studying humanity through the scientific method. Cultural anthropologists rely on the “softer” methods of
observation, participation, and interviews. Someone who studies the genetic distribution of blood types and
someone who studies an addiction treatment program may have a difficult time finding common ground.

Increasingly, however, urgent concerns such as inequality and climate change have highlighted the
importance of an integrated approach to the study of humanity. The issue of racial inequality is an excellent
example. Beginning with an approach from the cultural side of our discipline, many anthropologists explore
what we think we know about the concept of race. How many racial categories do you think there are in the
world? How can you tell a person’s racial identity? What do you know about your own racial category?

Biological anthropologist Jada Benn Torres and cultural anthropologist Gabriel Torres Colón teamed up to
explore how people use genetic ancestry testing to construct notions of collective history and racial belonging
(2020). For instance, if you learn through genetic testing that your ancestors most likely came from Nigeria,
you might begin to feel a certain identification with that country and with the continent of Africa as a whole.
You might begin to feel that you have less in common with the people of your country of citizenship and more
in common with the people of your country of ancestry, a racial connection perhaps felt as more fundamental
than the sociocultural connection to your home culture. While concerned about the potential for spreading
misconceptions about racial categories, Torres and Colon also note that racialized solidarity across national
boundaries can foster transnational movements for social justice. Such research shows how we actively
18 1 • What Is Anthropology?

construct our concepts about race using biological information about ourselves, all the time believing that
those concepts are embedded in nature.

FIGURE 1.6 This map shows the predicted skin colors of people based on the levels of ultraviolet radiation in the
areas where they live. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

Importantly, biological anthropology demonstrates that our common notions of race are inaccurate. Biological
anthropologists such as Agustín Fuentes (2012) and Nina Jablonski (2006) have looked carefully at the global
distribution of human traits such as skin color, facial features, hair texture, and blood type, among other
markers, in order to determine if humans are indeed grouped into discrete categories based on race. Short
answer: biologically speaking, there are no real racial categories. Each human trait varies along a spectrum,
and the various traits are mixed and matched among people in ways that make racial distinctions impossibly
inaccurate. As an example, take the issue of skin color, which is the most common way people assign race.
Jablonski demonstrates that skin color varies along a spectrum, from pinkish beige to dark brown, with people
throughout the world having skin of every possible shade between those two. Originally, humans evolving on
the African continent had dark skin to protect them from the direct ultraviolet light of the sun. As some early
humans migrated north into environments with less direct sunlight, their skin lightened to allow the
absorption of vitamin D from the much weaker sunlight.

Today, if we look at people with deep historical connections to particular geographical areas, we find that skin
color shifts gradually with location. Imagine setting out on a road trip from Kinshasa, the capital of the
Democratic Republic of the Congo, just a few degrees south of the equator in central Africa, and traveling all the
way up to the city of Tromsø in Norway, north of the Arctic Circle. This 157-hour trip would take you through
Nigeria, Niger, Algeria, Spain, France, Germany, Denmark, and Sweden. If you were paying attention to the skin
color of the indigenous peoples in each location, you would notice a gradual shift from deep brown in Kinshasa
to lighter brown in Algeria to dark beige in southern Spain to lighter beige in Sweden. You might also notice
other changes, such as more green and blue eyes and more red and blond hair, as you head into northern
Europe. At no point in your trip could you identify a boundary between groups. Rather, you would see a gradual
spectrum of change.

Whether looking at visible characteristics such as skin color or invisible genetic markers such as blood type,
biological anthropologists have demonstrated time and time again that there is no scientifically justifiable way
to divide the human population into racial categories. Any way you draw the lines, there will be more variation
within categories than between categories.

Does this mean that race does not exist? In terms of biology, that is exactly what it means. But in terms of social
reality, unfortunately not. Race does not exist in nature, but race does exist in our minds, our practices, and
our institutions. Archaeological excavations of the material lives of various groups in the United States,
including people from China and Ireland as well as enslaved peoples from Africa, show how notions of race
shaped their whole ways of life: the buildings in which they lived, the clothing they wore, the property they

Access for free at openstax.org


1.3 • Overcoming Ethnocentrism 19

owned, and the structure of their families (Orser 2007; Singleton [1985] 2016). In contemporary societies,
cultural anthropologists studying forms of racial inequality in societies all over the world—including the
United States, the Dominican Republic, Brazil, Japan, Kenya, and Zimbabwe—have uncovered the different
ways that each of these societies constructs racial categories and uses various criteria to assign (and often
reassign) race to a particular person.

Moreover, in-depth ethnographies illuminate the severity of racism in the everyday lives of people of color in
the United States and elsewhere. After three years of fieldwork on the West Side of Chicago, anthropologist
Laurence Ralph documented the suffering of people in this Black neighborhood as they contend with
discrimination, economic deprivation, gang violence, and political marginalization (2014). Ralph emphasizes
that the people he observed dream of a better life for themselves and their children, in spite of these struggles,
and describes how many turn to social and political activism in an attempt to make their neighborhood a
better place for everyone who lives there.

Linguistic anthropologists are interested in how race is constructed and expressed through language.
Marcyliena Morgan studied the underground hip-hop scene in Los Angeles, exploring how Black emcees and
musicians craft linguistic codes that reference their experiences of police violence, urban unrest, gang activity,
and gentrification (2009). Like Ralph, Morgan highlights the creativity and resilience of Black American
communities in the face of enduring racism in American society.

Taken together, these various anthropological approaches to race provide more insight and understanding
than any one approach ever could. Overturning the biological myth of race is essential to understanding the
complex reality of human diversity, but it is not enough. It would be a mistake to pretend that racial categories
do not matter just because the concept of race has no basis in biology. The combined work of archaeologists,
cultural anthropologists, and linguistic anthropologists demonstrates how the mythic notion of race has been
used to exploit and marginalize certain people throughout history and into the present. We also see how
people respond to racial subjugation with creativity and resilience, inventing cultural forms of resistance and
mobilizing their communities through social activism.

1.3 Overcoming Ethnocentrism


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define the concept of ethnocentrism and explain the ubiquity of ethnocentrism as a consequence of
enculturation.
• Distinguish certain forms of ethnocentrism in terms of their historical relationship to forms of empire and
domination.
• Identify primitivism in European and American representations of African peoples.
• Identify orientalism in European and American representations of Asian and Middle Eastern peoples.

Have you ever known somebody who seems to think the world revolves around them? The kind of friend who is
always talking about themselves and never asks any questions about you and your life? The kind of person who
thinks their own ideas are cool and special and their own way of doing things is absolutely the best? You may
know the word used to describe that kind of person: egocentric. An egocentric person is entirely caught up in
their own perspective and does not seem to care much about the perspectives of others. It is good to feel proud
of your personal qualities and accomplishments, of course, but it is equally important to appreciate the
personal qualities and accomplishments of others as well.

The same sort of “centric” complex operates at the level of culture. Some people in some cultures are
convinced that their own ways of understanding the world and of doing things are absolutely the best and no
other ways are worth consideration. They imagine that the world would be a much better place if the superior
beliefs, values, and practices of their own culture were spread or imposed on everyone else in the world. This
is what we call ethnocentrism.

Enculturation and Ethnocentrism


We are all brought up in a particular culture with particular norms and values and ways of doing things. Our
20 1 • What Is Anthropology?

parents or guardians teach us how to behave in social situations, how to take care of our bodies, how to lead a
good life, and what we should value and think about. Our teachers, religious leaders, and bosses give us
instruction about our roles, responsibilities, and relationships in life. By the time we are in our late teens or
early twenties, we know a great deal about how our society works and our role in that society.

Anthropologists call this process of acquiring our particular culture enculturation. All humans go through this
process. It is natural to value the particular knowledge gained through our own process of enculturation
because we could not survive without it. It is natural to respect the instruction of our parents and teachers who
want us to do well in life. It is good to be proud of who we are and where we came from. However, just as
egocentrism is tiresome, it can be harmful for people to consider their own culture so superior that they
cannot appreciate the unique qualities and accomplishments of other cultures. When people are so convinced
that their own culture is more advanced, morally superior, efficient, or just plain better than any other culture,
we call that ethnocentrism. When people are ethnocentric, they do not value the perspectives of people from
other cultures, and they do not bother to learn about or consider other ways of doing things.

Beyond the sheer rudeness of ethnocentrism, the real problem emerges when the ethnocentrism of one group
causes them to harm, exploit, and dominate other groups. Historically, the ethnocentrism of Europeans and
Euro-Americans has been used to justify subjugation and violence against peoples from Africa, the Middle
East, Asia, and the Americas. In the quest to colonize territories in these geographical areas, Europeans
developed two main styles of ethnocentrism, styles that have dominated popular imagination over the past two
centuries. These styles each identify a cultural “self” as European and a cultural other as a stereotypical
member of a culture from a specific region of the world. Using both of these styles of ethnocentrism,
Europeans strategically crafted their own coherent self-identity in contrast to these distorted images of other
cultures.

Primitivism and Orientalism


Since the 18th century, views of Africans and Native Americans have been shaped by the obscuring lens of
primitivism. Identifying themselves as enlightened and civilized, Europeans came to define Africans as
ignorant savages, intellectually inferior and culturally backward. Nineteenth-century explorers such as Henry
M. Stanley described Africa as “the dark continent,” a place of wildness and depravity (Stanley 1878). Similarly,
European missionaries viewed Africans as simple heathens, steeped in sin and needing Christian redemption.
Elaborated in the writings of travelers and traders, primitivism depicts Africans and Native Americans as
exotic, simple, highly sexual, potentially violent, and closer to nature. Though both African and Native
American societies of the time were highly organized and well-structured, Europeans often viewed them as
chaotic and violent. An alternative version of primitivism depicts Africans and Native Americans as “noble
savages,” innocent and simple, living in peaceful communities in harmony with nature. While less overtly
insulting, the “noble savage” version of primitivism is still a racist stereotype, reinforcing the notion that non-
Western peoples are ignorant, backward, and isolated.

Europeans developed a somewhat different style of ethnocentrism toward people from the Middle East and
Asia, a style known as orientalism. As detailed by literary critic Edward Said (1979), orientalism portrays
peoples of Asia and the Middle East as irrational, fanatical, and out of control. The “oriental” cultures of East
Asia and Middle East are depicted as mystical and alluring. The emphasis here is less on biology and nature
and more on sensual and emotional excess. Middle Eastern societies are viewed not as lawless but as
tyrannical. Relations between men and women are deemed not just sexual but patriarchal and exploitative.
Said argues that this view of Asian and Middle Eastern societies was strategically crafted to demonstrate the
rationality, morality, and democracy of European societies by contrast.

In his critique of orientalism, Said points to the very common representation of Muslim and Middle Eastern
peoples in mainstream American movies as irrational and violent. In the very first minute of the 1992 Disney
film Aladdin, the theme song declares that Aladdin comes from “a faraway place / where the caravan camels
roam / where they cut off your ear if they don’t like your face / it’s barbaric, but hey, it’s home.” Facing criticism
by antidiscrimination groups, Disney was forced to change the lyrics for the home video release of the film
(Nittle 2021). Many thrillers such as the 1994 film True Lies, starring Arnold Schwarzenegger, cast Arabs as
America-hating villains scheming to plant bombs and take hostages. Arab women are frequently portrayed as

Access for free at openstax.org


1.4 • Western Bias in Our Assumptions about Humanity 21

sexualized belly dancers or silent, oppressed victims shrouded in veils. These forms of representation draw
from and reproduce orientalist stereotypes.

Both primitivism and orientalism were developed when Europeans were colonizing these parts of the world.
Primitivist views of Native Americans justified their subjugation and forced migration. In the next section,
we’ll explore how current versions of primitivism and orientalism persist in American culture, tracing the
harmful effects of these misrepresentations and the efforts of anthropologists to dismantle them.

1.4 Western Bias in Our Assumptions about Humanity


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define and recognize cultural bias.
• Analyze forms of cultural bias in our own interactions and institutions.
• Describe how the four fields of anthropology can work together to expose and overturn the misconceptions of
cultural bias.

Euro-American ethnocentrism is everywhere in American culture—in our movies, advertising, museums,


amusement parks, and news media. Though the styles have shifted somewhat in the past century, both
primitivism and orientalism still persist as two discernible styles of bias.

Primitivism and Orientalism in Popular Culture


Think for a minute about the last time you saw an image of an African person. Was it, perhaps, an image of
wide-eyed girl in tattered clothing in an advertisement from a development agency requesting a charitable
donation? Or maybe it was a news media photograph of a child soldier wielding an AK-47 in a conflict zone in
the Democratic Republic of the Congo or another African country. Africa is still popularly represented as a dark
place full of deprivation and crisis. Africans are frequently infantilized as simple children who need the
support and tutelage of White Western helpers. But isn’t it true, you may say, that poverty and violent conflicts
are widespread in Africa? Isn’t the representation accurate to some degree?

The most troubled places on the African continent are the places where European colonialism was most brutal
and violent. In what is now the Democratic Republic of the Congo, the Belgian king Leopold II oversaw a reign
of terror against the local peoples, encouraging their enslavement for the lucrative rubber trade. Elsewhere in
Africa, European colonial governments stole land from local peoples and confined them to reservations,
forcing them to work on European plantations in order to pay taxes to the colonial government. Colonial
officials fomented conflict by privileging some ethnic groups and repressing others. Where you see violence
and conflict in Africa today, the roots can often be traced to the colonial period. Is this painful history included
in American representations of Africa?

Moreover, there are many bright spots in Africa, places such as Ghana and Botswana, with growing economies
and stable democracies. Would it surprise you to learn that Ghana has a space program? That there are more
mobile phones than people in Kenya? That several electric cars are manufactured in Africa?

Similar distortions are applied to Native Americans, frequently represented as victims of history, poor and
helpless, in need of outside help. The primitivist gaze shapes the representation of Native Americans in
museums, which often feature dioramas of humble people with stone tools, buckskin clothes, and tepees,
either living a simple life close to nature or engaged in tribal warfare, their bodies painted with vibrant colors.
Of course, Native Americans do not live this way now, but these are the images that come to mind in the
popular imagination. It is of course important for non-Native Americans to learn about the cultures of Native
peoples before and during their contact with European settlers, but it is equally important to understand the
legacies of history in the contemporary living conditions and activities of Native communities. Rather than
seeing Native peoples as passive victims, popular culture should also depict the dynamic and creative
responses of Native Americans to the forms of cultural violence enacted against them.
22 1 • What Is Anthropology?

FIGURE 1.7 One example of a healthy Native American dish is Navajo mutton stew with blue corn and dry bread.
(credit: “Mutton Stew with Blue Corn and Dry Bread” by Neeta Lind/flickr, CC BY 2.0)

For instance, did you know that a Native food movement is surging across the United States, both on Native
reservations and in American cities? Native food activists such as Karlos Baca and Sean Sherman are reviving
and reinventing the balanced, healthy cuisines of their ancestors, featuring dishes such as braised elk leg and
maple red corn pudding. Sherman and his partner, Dana Thompson, have founded the nonprofit group North
American Traditional Indigenous Food Systems (NATIFS), devoted to preserving Native foodways. The group
offers opportunities for tribes to set up Native cuisine restaurants, providing jobs and profits to communities
with high unemployment. Watch this video (https://openstax.org/r/sheansherman) to learn more about Sean
Sherman and the Native Food movement.

Like primitivism, orientalism has endured in American and European cultures. In the two decades following
the al-Qaeda attacks on American targets on September 11, 2001, the most prominent example of orientalism
in American culture has been the stereotype that all Islamic peoples are fanatical and violent. The
indiscriminate application of this stereotype to Islamic peoples across the Middle East was a major contributor
to the 2003 American invasion of Iraq, a country that had nothing at all to do with the September 11 attacks. To
promote the invasion, politicians used the orientalist notion that Iraq was a violent and irrational country
stockpiling weapons of mass destruction (which turned out to be false). As the war raged on, the Iraqi people
came to be categorized as either “unlawful combatants” or helpless victims of a cruel dictator. American
officials argued that Iraqis needed the help of American troops to save them from their subjugation and teach
them democracy.

For many Europeans and Americans, these forms of ethnocentric bias distort views of peoples living in large
geographical regions of the globe. Misunderstanding other cultures this way can result in policies and military
actions that do not achieve desired results. Moreover, ethnocentric bias promotes and reinforces inequality
among social groups within multicultural societies. When people with certain ethnic or racial identities are
seen as helpless or violent, they face discrimination in their pursuit of education, employment, and justice.

The Bias of Backwardness


Common to both primitivism and orientalism is the notion that European and Euro-American cultures are
more advanced and civilized than other cultures. Since at least the 19th century, Euro-American thinking has
been dominated by the idea that the various cultures of the world can be evaluated on a scale of sociocultural
sophistication from least advanced to most advanced. Typically, Native American and African cultures were
considered the most primitive, while those of Asia and the Middle East were thought of as slightly more
developed but certainly not as civilized as the societies of Europe, which were ranked at the top as the epitome
of human progress.

Access for free at openstax.org


1.4 • Western Bias in Our Assumptions about Humanity 23

Early anthropology played a role in promoting this ethnocentric way of thinking. Nineteenth-century
anthropologists detailed various hypothetical schemes charting the developmental stages that each culture
would go through in its pursuit of the European ideal of civilization. One very prominent scheme was proposed
by the British anthropologist Edward Tylor. Tylor suggested that each culture progressed from “savagery” to
“barbarism” to “civilization.” Since the change from one stage to another could not be witnessed by the
researcher, such “evolutionary” schemes were largely based on hypothetical conjecture, sometimes called
“theorizing from the armchair.”

While some anthropologists played a role in popularizing this way of thinking, others worked to expose it as
misguided and inaccurate. The writings of American anthropologist Franz Boas highlighted the fact that no
culture is isolated in its process of developmental change. Instead, each culture develops through interactions
with other cultures, as new ideas and inventions diffuse from one culture to the next. Moreover, cultural
change is not structured by an overall trajectory of progress as defined by the European example; rather,
cultures change in many ways, sometimes adopting new ways of doing things and other times reviving and
reclaiming older ways. Through these varied patterns of change, each culture forges its own unique history.

While the evolutionary schemes of 19th-century anthropology have been disproven, the underlying notion of
sociocultural progress toward a Euro-American ideal is still a widespread form of ethnocentric bias outside of
anthropology. Many people still refer to some countries as “developed” and “modern” and others as
“undeveloped” and “backward.” Think for a minute: Which countries are generally thought of as modern?
Which ones are frequently referred to as undeveloped? What is really meant by these labels?

These labels are rooted in Euro-American values. Championing capitalism and technology, many Europeans
and Americans view the generation of material wealth as the primary measure of the success of any society.
The divide between the more and less “advanced” countries of the world is largely a distinction between the
richer and poorer countries. European and American societies, which have become wealthy through the
development of global trade and industrial capitalism, are considered the most successful. Societies that have
not achieved the levels of wealth and technology associated with Euro-American industrial capitalism are
sometimes labeled “undeveloped.” Societies that have not industrialized at all are sometimes called
“premodern” or simply “traditional.”

As with older evolutionary schemes, this way of thinking relies on the notion that each society pursues
economic development in isolation. The poorer countries of the world are told: if you work hard and apply the
correct economic policies, then you too can become rich like the United States, the United Kingdom, and
Germany. But how did those countries become rich in the first place? Certainly not in isolation. The Boasian
emphasis on cultural interaction also applies to economic change. To a large degree, European and American
societies became wealthy by dominating other societies and keeping them poor. European countries
constructed a system of global capitalism designed to make them very rich by extracting raw materials and
human labor from their colonies. In fact, that was the whole impetus for colonialism.

The cultural anthropologist Sidney Mintz is one of many who have studied how this happened. Mintz explored
how European merchants designed a very lucrative system of production and consumption based on sugar
(1985). As European consumers began developing a taste for sugar in the 17th century, European merchants
developed sugar plantations in the New World using the labor of enslaved people transported from West
Africa. Sugar produced on these plantations was exported to Europe and the rest of the world, earning a hefty
profit for the European merchants who designed the system. Local people living in the places where sugar was
produced did not benefit much from this trade, and enslaved people suffered and died for it. Similar systems
were developed for the production of other global commodities such as cocoa, coffee, tea, and cotton. Some
commodities required enslaved labor and others involved small farmers, but the basic structure of the trade
was the same. The economies of many South Asian and African countries were designed entirely around the
export of primary commodities, the production of which was controlled by European merchants who reaped
the profits from this global trade. Many postcolonial countries still rely on the export of these primary
commodities.

What do these historical processes mean for understanding the world today? European merchants and
governments crafted strategic ways of thinking about the parts of the world they wanted to invade and
24 1 • What Is Anthropology?

colonize. To justify the development of the slave trade, the plantation system, and colonial rule, Europeans
labeled many non-Europeans as backward peoples needing the civilizing influence of European domination.
This form of bias persists in contemporary notions of backwardness applied to the poorer peoples and parts of
the world.

In reality, the colonial system was a global mechanism for European merchants and governments to extract
wealth from other parts of the world. European merchants took great care to maintain control over these forms
of highly profitable trade, edging out local merchants and forbidding local competition. Even today, we see the
remnants of this system in Euro-American domination of global trade. If the world seems divided between rich
and poor, it is not because some countries work hard and others are “backward.” It is because the global
system was founded on forms of inequality that endure into the present.

PROFILES IN ANTHROPOLOGY

Franz Boas
1858–1942

FIGURE 1.8 Franz Boas (credit: “FranzBoas” by Canadian Museum of History/Wikimedia Commons, Public
Domain)

Personal History: Franz Uri Boas was born in Germany to a middle-class Jewish family (Peregrine 2018). After
completing a PhD in physics and mathematics, he worked as a geographer on an expedition to the Canadian
Arctic, living and working with the Native Inuit peoples on Baffin Island. With his newfound passion for Native
American culture, Boas returned to Germany to work at a museum and began conducting ethnographic and
linguistic research among Native groups. In 1887, he came to the United States and established the first
anthropology department at Clark University in Massachusetts. He spent most of his career as an anthropology
professor at Columbia University and curator at the American Museum of Natural History in New York City.

Areas of Anthropology: Though he promoted a holistic approach integrating the four fields of anthropology,
Boas was primarily a cultural anthropologist specializing in the Native peoples of the Northwest coast of North
America. Between 1886 and 1900, he conducted 29 months of fieldwork in the region, focusing on the

Access for free at openstax.org


1.5 • Holism, Anthropology’s Distinctive Approach 25

Kwakiutl peoples of Vancouver Island. He recorded myths, songs, and folklore in Native languages and
described cultural activities such as food collection and artistic styles. Focusing on the linguistic and
psychological aspects of this rich ethnographic data, Boas sought to understand Native perspectives and
values. As the leading anthropologist of his time, he established an American tradition of recording
ethnographic observations in meticulous detail and promoted the goal of reaching for an insider’s point of
view.

Accomplishments in the Field: Boas profoundly disagreed with ethnocentric and racist theories circulating in
the social sciences in the late 19th and early 20th centuries. Some anthropologists of the day identified some
cultures as “primitive” or “savage,” arguing that each culture developed in isolation along a common trajectory
toward “civilization.” Rejecting this model, Boas used his ethnographic data to show that cultures do not
develop in isolation toward a common goal. Rather, each culture has its own unique historical trajectory, and
cultures are constantly changing by sharing new ideas and practices.

Importance of His Work: Boas was horrified by the use of anthropological methods to support the theories
and practices of White supremacy. In the 19th century, some American researchers measured the skulls of
various ethnic groups, arguing that people who had immigrated to the United States from northern Europe had
larger skulls and were therefore intellectually superior. In 1907, Boas conducted a survey for the U.S.
Immigration Commission measuring the skulls of 17,821 American immigrants and their children. Comparing
the head shapes of parents and children, Boas discovered that the children had larger skulls due to
environmental factors in their new homeland, such as diet and medical care. His findings dealt a strong blow
to race theory. Throughout his career, Boas spoke out against racism, arguing that biological differences have
nothing to do with culture, language, or achievement.

1.5 Holism, Anthropology’s Distinctive Approach


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define and give examples of holism.
• Analyze how different elements of society cohere with and reinforce one another.
• Identify how different elements of society can contradict one another, motivating social change.

In 2020, the COVID-19 pandemic swept across the globe. Nearly 210 million people had fallen sick with the
coronavirus and more than 4 million had died as of August 2021. Medical researchers are still studying the
long-term effects of this illness on the lungs and brains of people who have recovered. Some have discovered
psychological effects as well, such as increased risks for depression, anxiety, and schizophrenia.

Beyond the medical realm, the effects of the pandemic reached into every aspect of our societies and our
everyday lives. In societies all over the world, people were forced to remain at home, “sheltering in place” from
the dangers of the disease. Businesses closed their doors to the public, and many shut down permanently,
unable to pay their bills. By May 2020, nearly 50 million Americans had reported losing their jobs due to the
pandemic. The epidemic of disease ballooned into an epidemic of grief as people mourned the loss of the those
who had died and worried about those who had fallen sick. Stressed out by so many disruptions, some adults
turned to alcohol and drugs, and addiction rates soared. Incidents of domestic violence escalated. Racial
violence against Asian Americans increased as some Americans blamed China for the emergence and global
spread of the disease. People everywhere reported feeling lonelier and more cut off from their friends and
family members.

And yet there were also some positive consequences. Because people were not driving as much, air quality
improved in many urban areas, giving relief to many people who suffer from asthma. Looking up into the night
sky, some people were able to see stars for the very first time. Some people reported valuing their friends and
family members even more now that they could not spend time with them in person. New social media
technologies spread, such as Zoom, and many people learned to use existing technologies such as FaceTime
and Skype. People also became aware of the valuable contributions made by “essential workers” in drugstores,
hardware stores, and grocery stores as well as hospitals and nursing homes.
26 1 • What Is Anthropology?

How did a virus cause so many changes? The various elements of society are entwined in a complex whole.
Dramatic changes in one area, such as epidemic disease in the realm of public health, can trigger a chain of
effects throughout other social realms, such as the family, the economy, religion, and the political system.

You’ll recall the word holism from our earlier discussion about anthropology’s commitment to understanding
how the many parts of society work together. Holism is a distinctive method of analysis that foregrounds the
ever-changing relationships among different realms of culture.

Society as an Integrated Whole


Throughout the 2010s, infant death rates in certain rural areas in Africa decreased dramatically. While thrilled
with this positive trend, researchers did not initially know how to explain it. Were mothers and fathers doing
something different to promote the health of their babies? Were African governments providing better health
services for infants? Were aid agencies providing more resources? None of these things seemed to be true in
any significant way.

The one thing that had changed in the areas with lower infant mortality was the spread of mobile phones.
Could that have something to do with lower infant mortality? And if so, how? Researchers hypothesize that it
wasn’t just the possession or use of mobile phones that was making the difference—it was the capability to use
mobile money transfers and other fintech. If a baby had a fever in the middle of the night, the mother could
now immediately text members of her extended family to organize the necessary funds to take the baby to a
hospital for treatment. Quicker treatment meant a better chance for recovery. Something that does not appear
to be directly related to infant health may in fact have a great impact on it.

Recall from the beginning of this chapter our discussion of the very broad scope of anthropology. While other
disciplines focus on one realm of society, such as medicine or technology, anthropology ranges across all
realms of human thought and activity. Using the technique of holism, anthropologists ask how seemingly
disparate elements of social life might be related in unexpected ways.

In American and European cultures, the most common form of marriage is a union of two people. In the United
States, many marriages end in divorce and most people then remarry, resulting in a cycle of marriage-divorce-
remarriage called serial monogamy. In other cultures, however, a man may have more than one wife. It might
be tempting to think that the dominant form of marriage in a culture is related to morality or gender relations.
It turns out, however, that one very significant influence on marriage patterns is the food-getting strategy of a
particular culture. In small-scale farming cultures, the marriage of one man to two or more women provides an
abundance of children to help out with the work of weeding, watering, fertilizing, and guarding the crops
(Boserup [1970] 2007; Goody 1976). In cultures where children contribute to food production, the marriage of
one man to multiple women is more prevalent. This isn’t always the case, of course, as there are other factors
that influence the form of marriage practiced in a culture, but the useful work of children does contribute to
the popularity of this form of marriage.

In the contemporary United States, by contrast, most people work not on farms but in offices, shops, and
factories. Children are not valued as sources of household labor, and they are not legally permitted to work for
wages. In fact, children can be viewed as a drain on the household, each one requiring a massive investment of
resources in the form of health care, childcare, special equipment, educational opportunities, and expensive
toys. In this context, the increased fertility of multiple wives might impoverish the household. Moreover, our
fast-paced, capitalist economy requires a flexible and highly mobile work force. American workers can lose
their jobs, and they must be prepared to move and retrain in order to find further work. Many Americans
experience periods of uncertainty and precarity in their work lives, conditions that affect the livelihood of their
households as well as their relationships with their marriage partners and children. Such a context contributes
to smaller family size and fragile marriage bonds. The cycles of stability and disruption in American work life
are mirrored in the cycles of marriage and divorce involved in serial monogamy.

These are just two examples of why anthropologists are committed to taking such a broad view of the cultures
they study. Often, the various realms of society are related in ways that are not at first apparent to the
researcher. By specializing too narrowly on only one realm, the researcher might miss the wider forces that
shape the object of study.

Access for free at openstax.org


1.5 • Holism, Anthropology’s Distinctive Approach 27

Sources of Contradiction, Conflict, and Change


Holistic analysis considers not only how the various features of culture hold together but also how change in
one feature can generate cascading changes among others. Often, anthropologists begin their analysis by
focusing on one significant change in the lives of a particular cultural group and then chart the ramifications of
that change through various other realms of culture.

Attiya Ahmad conducted research among South Asian women who migrate to the Middle East for jobs as
housekeepers (2017). She writes about how these women adapt to a new culture and living situation in Kuwait
and the disruptions they face when they return to their families and home cultures. On the job in Kuwait, these
domestic workers must learn to speak Arabic, operate household gadgets, prepare an entirely different
cuisine, respect Islamic norms and practices, and perform their appropriate gender role as female members of
a Kuwaiti household. They face the cultural requirement that women should be naram, or soft and malleable,
as they develop emotionally charged relationships with the various members of the household. These
requirements bring about profound personal transformations for these women as they deal with the
contradictions of being both successful wage earners and subordinated cultural others.

The motivation to migrate is primarily financial: the need to pay for schooling, marriages, medical care, and
other family expenses. While the women are working in Kuwait, their families become economically
dependent on the money they send back home even as their emotional relationships with their family
members become weaker and more difficult. When they return home, profoundly changed by their
experiences in Kuwait, their natal families nonetheless expect them to behave exactly as they did before they
left, observing the same gender and age-related norms that govern the household. This creates a sense of
internal conflict for these women. Unable to truly reintegrate with their natal families, many either seek out
new connections in their home communities or migrate back to Kuwait. Some begin learning more about
Islam by attending special da’wa classes, where they meet other women in the same situation. Finding ethical
inspiration in Islamic teachings, many do convert, against the objections of their natal families and their
Kuwaiti employers.

All cultures are constantly changing, with small changes in one realm snowballing into larger and larger
changes within and beyond that culture. The Me Too movement is another good example. What began in 2006
as a call by American activist Tarana Burke for solidarity and empathy with victims of sexual harassment has
now spread into many sectors of American society and across the globe. Initially focused on high-profile
celebrities and the movie industry, the Me Too movement has raised awareness of widespread sexual
harassment and assault in the fashion industry, churches, the finance industry, sports, medicine, politics, and
the military. Activists press for legal changes to protect workers, especially whistleblowers who come forward
with allegations of inappropriate sexual behavior. Evaluations of patriarchal and chauvinistic behavior in these
institutional realms have sparked scrutiny of the more informal cultural norms of American romance and
dating. The Me Too movement challenges the way Americans think about the gender roles of men and women,
appropriate speech and gestures, and the distinction between public life and private life.

The movement has prompted processes of dialogue and change in at least 28 other countries, including
Afghanistan, China, Nigeria, and the Philippines. The global campaign has been interpreted differently in each
of these cultural contexts as the transcultural intentions of American activists intersect with local norms of
gender and sexuality. Indeed, some critique the Me Too movement as ethnocentric. Though the calls for reform
resonated with French feminists, Me Too activism sparked a backlash among many other French people, with
some men and even women arguing that French men should have the right to make sexually provocative
comments and rub against women in public places.

While many anthropologists actively support the Me Too movement, our methods of cross-cultural
comparison call on us to set aside our personal values (at least temporarily) in order to understand how people
in various cultural contexts interpret and act on the cross-cultural campaign against gender-based
harassment and assault. This method of suspending personal values is key to understanding how all the
elements of a particular culture interact with one another, including pressures from the outside.
28 1 • What Is Anthropology?

1.6 Cross-Cultural Comparison and Cultural Relativism


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define the concept of relativism and explain why this term is so important to the study of anthropology.
• Distinguish relativism from the “anything goes” approach to culture.
• Describe how relativism can enlighten our approach to social problems.

Recall our earlier discussion of cultural styles of clothing. American clothing style is related to American
values. Ghanaian clothing style is related to Ghanaian values. We have seen how different realms of culture are
interrelated, fitting together to form distinctive wholes. Anthropologists use the term cultural relativism to
describe how every element of culture must be understood within the broader whole of that culture. Relativism
highlights how each belief or practice is related to all of the other beliefs and practices in a culture. The
anthropological commitment to relativism means that anthropologists do not judge the merits of particular
beliefs and practices but rather seek to understand the wider contexts that produce and reinforce those
elements of culture. Even when studying controversial topics such as piracy and guerilla warfare,
anthropologists set aside their personal convictions in order to explore the complex web of cultural forces that
determine why we do the things we do.

Relativism Is Not “Anything Goes”


Critics of the notion of relativism, believing so strongly in their own cultural norms that they cannot set them
aside, even temporarily. They argue that relativism is amoral, a refusal to condemn aspects of culture
considered to be wrong and harmful. For them, relativism means “anything goes.”

For anthropologists, cultural relativism is a rigorous mode of holistic analysis requiring the temporary
suspension of judgment for the purposes of exploration and analysis. Anthropologists do not think that violent
or exploitative cultural practices are just fine, but they do think that the reasons for those practices are a lot
more complex than we might imagine. And frequently, we find that the judgmental interventions of
ethnocentric outsiders can do more harm than good.

Morality, Activism, and Cultural Relativism


A striking example of the application of cultural relativism in anthropology is the controversy surrounding
female genital cutting (FGC), sometimes called female genital mutilation. FGC is a cultural practice in which an
elder cuts a younger woman’s genitalia, removing all or part of the clitoris and labia. The practice is common in
parts of Africa and the Middle East. FGC is not only extremely painful; it can also lead to infection, urination
problems, infertility, and complications in childbirth.

The World Health Organization and the United Nations condemn the practice as a form of violence against
children, a danger to women’s health, and a violation of basic human rights. These organizations view FGC as a
form of discrimination against women, enforcing extreme inequality among the sexes. Efforts to ban FGC have
focused on educating parents and children about the medical harms associated with the practice. Local
governments are encouraged to enact laws banning FGC and impose criminal penalties against the elders who
perform it.

Access for free at openstax.org


1.6 • Cross-Cultural Comparison and Cultural Relativism 29

FIGURE 1.9 Rendille Kenyan women attending a church dedication ceremony. (credit:
“180818_TSCOKenya_EstherHavens_0997” by Ann/flickr, CC BY 2.0)

Despite decades of campaigning against FGC, however, the practice remains widespread. If condemning FGC
has not been effective in reducing it, then what can be done? Anthropologist Bettina Shell-Duncan has taken a
more relativist approach, attempting to understand the larger cultural norms and values that make FGC such
an enduring practice. Setting aside her personal opinions, Shell-Duncan spent long periods in African
communities where FGC is practiced, talking to people about why FGC is important to them. She learned that
FGC has different functions in different sociocultural contexts. Among the Rendille people of northern Kenya,
many people believe that men’s and women’s bodies are naturally androgynous, a mix of masculine and
feminine parts. In order for a girl to become a woman, it is necessary to remove the parts of female genitalia
that resemble a man’s penis. Likewise, in order for a boy to become a man, the foreskin must be removed
because it resembles the folds of female genitalia.

Other societies value FGC for different reasons. Some Muslim societies consider FGC a form of hygiene,
making a girl clean so that she can pray to Allah. Some communities see FGC as a way of limiting premarital
sex and discouraging extramarital affairs. In the colonial period, when FGC was banned by the colonial
government, some Kenyan girls practiced FGC on themselves as a form of resistance to colonial authority. As
FGC is promoted and carried out by senior women in most contexts, the practice becomes a way for senior
women to solidify power and exert influence in the community.

People in communities practicing FGC are often aware of the efforts of outside groups to ban the practice. They
know about medical complications such as the risk of infection. But the denunciations of outsiders often seem
unconvincing to them, as those denunciations tend to ignore the cultural reasons for the endurance of FGC.
People who practice FGC do not do it because they despise women or want to harm children. Shell-Duncan
argues that parents weigh the risks and benefits of FGC, often deciding that the procedure is in the best
interest of their child’s future.

Personally, Shell-Duncan remains critical of FGC and works on a project with the Population Council designed
to dramatically reduce the practice. Cultural relativism does not mean permanently abandoning our own value
systems. Instead, it asks us to set aside the norms and values of our own culture for a while in order to fully
understand controversial practices in other cultures. By suspending judgment, Shell-Duncan was able to learn
two important things. First, while campaigns to eradicate FGC frequently target mothers, providing them with
educational material about the medical risks involved, Shell-Duncan learned that the decision to go ahead with
the procedure is not made by parents alone. A large network of relatives and friends may pressure a girl’s
parents to arrange for the cutting in order to ensure the girl’s chastity, marriageability, and fertility. Secondly,
Shell-Duncan learned that people who practice FGC do it because they want the best for their girls. They want
their girls to be respected and admired, considered clean and beautiful, fit for marriage and childbearing.

Shell-Duncan argues that outside organizations should reconsider their efforts, focusing more on
30 1 • What Is Anthropology?

communities than on individual parents. Awareness campaigns will be more effective if they resonate with
local norms and values rather than dismissively condemning them as part of the whole culture of FGC. Some
researchers urge anti-FGC activists to connect with local feminists and women’s groups in an effort to
empower local women and localize the movement against FCG. Some alternative approaches press for more
incremental forms of change, such as moving the practice to more sanitary conditions in clinics and hospitals
and reducing the severity of the procedure to smaller cuts or more symbolic nicks.

As this example illustrates, cultural relativism is not an amoral “anything goes” approach but rather a strategy
for forming cross-cultural relationships and gaining deeper understanding. Once this foundation has been
established, anthropologists are often able to revise their activist goals and more effectively work together with
people from another culture in pursuit of common interests.

1.7 Reaching for an Insider’s Point of View


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define the notion of insider’s point of view.
• Critique the notion of insider’s point of view, explaining how it is never perfectly achievable.
• List and describe the distinctive methods anthropologists deploy in their attempts to represent an insider’s
point of view

Bettina Shell-Duncan’s work on FGC demonstrates the importance of setting aside your own values and
opinions in order to see an issue from the point of view of those directly involved. This often means working
across contexts, whether studying another group or another culture. Anthropologists across the four fields
apply this technique. Cultural anthropologists talk to people and participate in social activities in order to
understand cultural life. Archaeologists rely on artifacts and fossils to reconstruct the sociocultural life of
peoples in earlier times and different places. Through these different methods, anthropologists all aim for the
same thing: they want to understand the perspectives of the people who practice a particular culture,
sometimes called an insider’s point of view.

The Challenge of Representing Others


The anthropological goal of representing an insider’s point of view is controversial. Is it truly possible to step
outside your own identity to really understand a different perspective? How can a researcher from a particular
culture possibly understand exactly how it feels to be a member of another culture? Even anthropologists who
study their own cultures may find themselves researching people from different classes, ethnicities, or gender
categories. Is it possible to accurately represent the perspectives of people whose lives are so different from
your own? Is it ethical? Is it valuable?

For decades, White European and American anthropologists conducted research and wrote ethnographies as
if the challenge of representing cultures very different from their own was really no problem at all. Empowered
by White privilege and ethnocentrism, many earlier anthropologists believed that long-term intensive
fieldwork was enough to give them cross-cultural insight into the perspectives of the people they studied.

Too frequently, those anthropologists reduced the complexity of the non-Western cultures they studied to just
one point of view, as if the people in that society all interpreted their cultural rules the same way and never
disagreed or changed the rules over time. In her book about Japanese culture, The Chrysanthemum and the
Sword (1946), anthropologist Ruth Benedict describes Japanese people in terms of common personality traits,
such as reverence for the emperor and a moral sense guided by shame. Critics have argued that her
conclusions are skewed by her overreliance on very few informants, all of them Japanese people confined to
internment camps during World War II. As we have explored in this chapter, every culture comprises multiple
perspectives that often contradict one another, generating sociocultural conflict and change. Recognizing this
situation, contemporary anthropologists often conduct research among several different subgroups and
geographical locations, integrating insights from these various arenas into a comprehensive and dynamic view
of cultural complexity.

Then there is the question of deep-seated bias, often operating unconsciously among researchers and the

Access for free at openstax.org


1.7 • Reaching for an Insider’s Point of View 31

people they study. Consider the situation above in which a White American anthropologist conducts research
in an African country previously colonized by Europeans. European colonialism left behind a legacy of White
privilege in postcolonial African countries. Earlier anthropologists did not often recognize how racialized
power dynamics might shape their research and writing, distorting their representations of the peoples they
studied. In the 1960s, anthropologists began to think more carefully about these issues, realizing that an
insider’s point of view is never perfectly achievable. As human beings, our own perspectives are conditioned
by our own enculturation, our own ways of seeing and thinking about the world around us.

If an insider’s point of view is never really possible, should we give up on this aspirational goal of the
discipline? In such a scenario, researchers would only study and write about people from the same
sociocultural categories as themselves. So, for example, Americans would only research and write about other
Americans. But are all Americans really members of the same sociocultural category? Could an upper-class
Asian American from Manhattan research and write about a poor Black community in the Deep South? Could a
Latino man write about a group of Latinx/Latina/Latino people consisting of all genders? American culture is
not unique in its complex array of identities. In all cultures, people have multiple identities as members of
multiple sociocultural categories. While you may be an insider within your culture in some respect, you may
be an outsider by some other measure. The ethical question of who can represent who is riddled with
difficulties.

Moreover, resigning ourselves to studying “our own people,” whoever they might be, is tantamount to giving up
on cross-cultural research and the insight, empathy, dialogue, and transformation that frequently result from
it. Anthropological insights have been key to rethinking American notions of sexuality, family, and race, among
so many other pressing issues. We need the skills of cross-cultural research now more than ever. While perfect
representations of different communities and cultures may be impossible, many anthropologists now deploy
innovative methods designed to address the problems of history and power at the heart of the discipline. The
aim is not to achieve perfect ethnography but to work ethically and collaboratively to produce what
contemporary cultural anthropologist Nancy Scheper-Hughes has termed “good enough ethnography.”

Collaborative Methods of Representation


Faced with the challenges of representation, many anthropologists practice methods of collaboration with the
individuals and groups that they study. Collaborative ethnography has a very long history in cultural
anthropology, traceable all the way back to early Euro-American ethnographies of Native Americans. Often,
anthropologists began their research by employing a local person as a translator or field assistant, a role that
usually evolved into something much more cooperative.
32 1 • What Is Anthropology?

FIGURE 1.10 Francis La Flesche (credit: “Francis laflesche” by National Anthropological Archives, Smithsonian
Institution/Wikimedia Commons, Public Domain)

Researching the Omaha peoples in the early 20th century, anthropologist Alice Cunningham Fletcher began
working with a young Omaha man, Francis La Flesche. Through their collaboration, La Flesche became an
ethnographer himself. While most anthropologists of the day merely acknowledged their local collaborators (if
they did even that), La Flesche became a full coauthor of their joint ethnography, The Omaha Tribe (1911).

Today, anthropologists collaborate with the people they study in a number of ways. Some involve local people
as readers and editors of their work, sometimes including community responses in the published
ethnography. Some conduct focus groups to generate local feedback on particular chapters. Some
anthropologists hold community meetings or forums to talk about the major themes and implications of their
work. And some, like Fletcher, collaborate with members of the local community as equal coauthors on books
and articles. Such methods strengthen ethnography by ensuring accuracy, promoting multiple perspectives,
and striving to make anthropological work more relevant to the communities being studied.

Collaboration also draws attention to the personal side of ethnography. Instead of extracting ethnographic
“facts” from the process of fieldwork, many contemporary anthropologists focus on describing particular
people, insightful conversations, and cooperative practices encountered in their research. Through this kind of
representation, culture is represented as a constellation of personal perspectives, each one shaped by the
position of each person in that community. Anthropologists also now acknowledge that ethnography is shaped
by the personal background and identity of the researcher as well as the motivations and intended audience of
the research. Collaborative anthropologists frequently describe their research in the first person, openly
acknowledging how their personal and cultural biases influence their research.

Anthropologist Luke E. Lassiter takes a collaborative approach in his study of the song and dance of
contemporary Kiowa communities of southern Oklahoma (1998). Lassiter describes how he became interested
in Kiowa song as a boy through his involvement in the Order of the Arrow, an affiliate of the Boy Scouts. Moving
beyond the superficial representations of Native American culture in Boy Scout teachings, Lassiter went on to
attend powwows, where he met singers and learned more about Kiowa culture. He developed a close
friendship with renowned Kiowa singer Billy Evans Horse, who taught Lassiter how to sing Kiowa songs and
encouraged him to pursue his interest in Kiowa culture in graduate school. Instead of foregrounding his own

Access for free at openstax.org


1.7 • Reaching for an Insider’s Point of View 33

description of Kiowa song and dance, Lassiter highlights the individual experiences and opinions of his local
collaborators as they describe how songs are created, passed down, and interpreted in the community.

Collaborative anthropology is not only more ethical and accurate; it is also more socially conscious and
political. When anthropologists collaborate as equals, they often become socially involved and politically
committed to the welfare of the communities they study. There are various terms for this, among them
engaged anthropology, public anthropology, anthropological advocacy, and applied anthropology. When those
communities face struggles over land, food security, medical care, or human rights abuses, many
anthropologists support their interests in a number of ways. Anthropologists often speak out publicly, write
sympathetic ethnographies, testify in court, participate in protests, and coordinate with organizations that can
provide material aid. Anthropologist Stuart Kirsch was researching magic and sorcery in a Yonggom village in
Papua New Guinea when he became concerned about pollution from local copper and gold mines nearby
(2018). As the community he was studying mobilized to protect their environment, Kirsch became involved in
their lawsuit against the Australian owners of the mine. He contributed to a social and environmental impact
study and advised lawyers representing the affected communities. He spoke out to local media and scholarly
publications, explaining the environmental problems caused by pollution from the mine.

Working across Cultures toward Common Goals


Stepping back for a moment, consider the problems facing us as humans on our shared planet. Climate change
threatens the survival of humanity and the biodiversity of plants and animals. Forms of deeply entrenched
inequality fuel racial, ethnic, and class conflicts within and between nations. These are global problems,
transnational problems, cross-cultural problems. Human beings need to find a way to communicate and
cooperate across the sociocultural boundaries that divide us, always recognizing the power dynamics involved
in that process.

How can we do this? Anthropology teaches us that we may never understand exactly how it feels to be a
member of a different culture or group within our own culture. But if we want to work together with people of
different sociocultural backgrounds to solve these pressing global issues, we have to try. Long-term fieldwork
and cross-cultural collaboration are not perfect solutions to the challenges of cross-cultural understanding,
but these methods give us a place to begin. And anthropological methods and insights can be transformative,
making possible the kinds of empathy and dialogue necessary to solve our global problems.

The goal of this anthropology textbook is to guide you in this process of transformation as you learn about the
cultural lives of the various peoples with whom you share this planet.

MINI-FIELDWORK ACTIVITY

Representation and Otherness

List three characters from fictional movies or television shows who represent people from cultures different
from your own. What adjectives would you use to describe these characters? How are they made to appear?
How do they act? Are they central or marginal characters? What role does each play in the plot or theme? What
might be the consequences of representing cultural groups in this way? Do you see evidence of ethnocentrism,
primitivism, and/or orientalism as described in this chapter?

Suggested Readings
Engelke, Matthew. 2018. How to Think Like an Anthropologist. Princeton: Princeton University Press.

Hastrup, Kirsten, ed. 2014. Anthropology and Nature. Routledge Studies in Anthropology 14. New York:
Routledge.

Otto, Ton, and Nils Bubandt, eds. 2010. Experiments in Holism: Theory and Practice in Contemporary
Anthropology. Malden, MA: Wiley-Blackwell.
34 1 • Key Terms

Key Terms
Anthropocene the contemporary period of so superior that no other culture is worth
increasing human impact on the ecosystems of consideration. Ethnocentric people often imagine
our planet. that the world would be a much better place if the
anthropology the study of humanity across time beliefs, values, and practices of their own culture
and space. were spread to or imposed on everyone else in
archaeoastronomy the study of how people in the the world.
past understood and used celestial objects for ethnography a written book or article about a
navigation, calendars, politics, and the timing of particular culture.
ritual events. excavation the removal of fossils and artifacts
archaeology the field of anthropology that relies from the ground in order to learn as much as
on the excavation of artifacts and fossils to possible about how people lived in times before
explore how environmental and historical and after the development of writing.
conditions have produced a diversity of human fieldwork a research method that requires cultural
cultures. anthropologists to live for many months or years
artifacts objects made by humans, such as pottery in the societies they study, adopting local ways of
or tools. living, eating, dressing, and speaking as closely
biological anthropology the field of anthropology as possible.
that focuses on the earliest processes in the forensic anthropology the application of the
biological and sociocultural development of techniques of biological anthropology to solve
human beings as well as the biological diversity of crimes.
contemporary humans. Biological fossils the remains of organism preserved in the
anthropologists study the origins, evolution, and environment.
diversity of our species. holism how the elements of human life are bound
cultural anthropology the field of anthropology together to form distinctive cultures.
devoted to describing and understanding the hominins the evolutionary ancestors of modern
wide variety of human cultures. Cultural humans.
anthropologists focus on such things as social insider’s point of view a goal of anthropological
thought, action, ritual, values, and institutions. research, representing the perspectives of people
cultural other a stereotype of a person from a who practice a particular culture.
different culture, used to create a cultural linguistic anthropology the field of anthropology
distinction between “us” and “them.” that explores the central role of language in
cultural relativism understanding every element human cultural life. Linguistic anthropologists
of culture within the broader whole of that study the origins of language, how language
culture. Cultural relativism highlights how each shapes thought, and how language operates as a
belief or practice is related to all of the other tool of power.
beliefs and practices in a culture. orientalism the depiction of some cultural groups,
enculturation the process of learning and particularly people from the Middle East and
acquiring a particular culture, often intensified in Asia, as exotic, irrational, fanatical, and sensuous.
childhood. primitivism the depiction of some cultural groups,
ethnobotany the study of how people in different particularly Africans and Native Americans, as
cultures categorize and use plants for food, exotic, simple, highly sexual, potentially violent,
shelter, tools, transportation, art, and religion. and closer to nature.
ethnocentrism the notion that one’s own culture is

Summary
Anthropology is an incredibly broad discipline, worked together in a wide variety of environmental
covering the entire scope of human experience, but and historical conditions to produce a diversity of
its enormity is controlled by a common narrative cultures. The three central commitments are
and set of three central commitments. The common exploring sociocultural diversity, examining how
narrative states that human beings have developed societies hold together, and studying the
flexible biological and social features that have interdependence of humans and nature.

Access for free at openstax.org


1 • Critical Thinking Questions 35

Anthropologists have developed four main anthropologists use holistic techniques of


approaches to pursuing anthropology’s common examination and analysis, seeking to understand
narrative, comprising the discipline’s four fields: how the various elements within a culture fit
biological anthropology, archaeology, cultural together and how these elements can contradict one
anthropology, and linguistic anthropology. Each of another, provoking change. Effective holistic
these fields generates a particular type of knowledge analysis requires a commitment to the method of
about the human experience that can be integrated cultural relativism, which requires a researcher to
with knowledge from the other three fields into a set aside their own personal values in order to
deeper, richer understanding of humanity’s central appreciate another culture on its own terms. An
challenges, such as racial injustice and climate important contribution to a rich appreciation of
change. another culture is the input and participation of
cultural insiders. The ethical challenges of
Getting at that deeper understanding,
understanding and representing another culture
anthropologists learn to recognize their own biases
have led anthropologists to develop collaborative
as forms of ethnocentrism such as primitivism and
ways of working with cultural insiders, aimed at
orientalism. Rather than categorizing societies
addressing the power asymmetries of fieldwork and
according to levels of sophistication (as European
ethnography.
scholars did in the 19th century), contemporary

Critical Thinking Questions


1. Have you ever taken a course in one of the other thinking about your own society?
social disciplines, such as economics, political 5. Identify a contemporary problem in your own
science, history, or religion? How would society. How would you pursue a holistic analysis
anthropology study the same subject matter in a of that problem? What are the various realms of
different way? culture that directly or indirectly relate to that
2. Which other social issues might benefit from a problem?
four-field approach? Propose one issue, and 6. Is it really possible to set aside your own personal
consider how each of the four fields might values when studying something you consider
contribute to our understanding of that issue. morally troubling or simply wrong? Identify a
3. Have you ever thought or said something controversial topic in your own or another
ethnocentric? What is an appropriate response if culture, ideally one that is personally meaningful
someone else says something ethnocentric in a to you. How would you practice cultural
conversation? How can people learn to recognize relativism when studying this topic? How would
and rethink ethnocentric notions? relativism change the way you interact with
4. As mentioned in this chapter, one very dominant people in the course of your research? How
way of evaluating the sophistication of different would it change the kinds of questions you would
societies is by measuring the amount of wealth ask in interviews?
generated by each one. Can you think of an 7. Make a list of possible ways you could collaborate
alternative way of evaluating progress or with someone from another social or cultural
development? Would that way reorder the global group in an effort to represent the perspectives of
hierarchy? How might it change your way of cultural insiders.

Bibliography
Ahmad, Attiya. 2017. Everyday Conversions: Islam, Domestic Work, and South Asian Migrant Women in
Kuwait. Durham, NC: Duke University Press.

Ball, Christopher. 2018. Exchanging Words: Language, Ritual, and Relationality in Brazil’s Xingu Indigenous
Park. Albuquerque: University of New Mexico Press.

Bayly, Susan. 1999. Caste, Society, and Politics in India from the Eighteenth Century to the Modern Age. The
New Cambridge History of India, vol. 4, no. 3. Cambridge: Cambridge University Press.

Bielo, James S. 2018. Ark Encounter: The Making of a Creationist Theme Park. New York: New York University
Press.
36 1 • Bibliography

Boserup, Esther. (1970) 2007. Woman’s Role in Economic Development. London: Earthscan.

Carr, E. Summerson. 2011. Scripting Addiction: The Politics of Therapeutic Talk and American Sobriety.
Princeton, NJ: Princeton University Press.

Dillehay, Tom D., ed. 2017. Where the Land Meets the Sea: Fourteen Millennia of Human History at Huaca
Prieta, Peru. Austin: University of Texas Press.

Edwards, David B. 2017. Caravan of Martyrs: Sacrifice and Suicide Bombing in Afghanistan. Oakland:
University of California Press.

Fredericks, Rosalind. 2018. Garbage Citizenship: Vital Infrastructures of Labor in Dakar, Senegal. Durham, NC:
Duke University Press.

Fuentes, Agustín. 2012. Race, Monogamy, and Other Lies They Told You: Busting Myths about Human Nature.
Berkeley: University of California Press.

Goody, Jack. 1976. Production and Reproduction: A Comparative Study of the Domestic Domain. Cambridge:
Cambridge University Press.

Haas, Randall, James Watson, Tammy Buonasera, John Southon, Jennifer C. Chen, Sarah Noe, Kevin Smith,
Carlos Viviano Llave, Jelmer Eerkens, and Glendon Parker. 2020. “Female Hunters of the Early Americas.”
Science Advances 6 (45): eabd0310. https://doi.org/10.1126/sciadv.abd0310.

Halvaksz, Jamon. 2006. “Drug Bodies: Relations with Substance in the Wau Bulolo Valley.” Oceania 76 (3):
235–244.

Heatherington, Tracey. 2010. Wild Sardinia: Indigeneity and the Global Dreamtimes of Environmentalism.
Seattle: University of Washington Press.

Isbell, Lynne A. 2009. The Fruit, the Tree, and the Serpent: Why We See So Well. Cambridge, MA: Harvard
University Press.

Jablonski, Nina G. 2006. Skin: A Natural History. Berkeley: University of California Press.

Keim, Curtis, and Carolyn Somerville. 2018. Mistaking Africa: Curiosities and Inventions of the American
Mind. 4th ed. New York: Routledge.

Kirsch, Stuart. 2018. Engaged Anthropology: Politics beyond the Text. Oakland: University of California Press.

Lassiter, Luke E. 1998. The Power of Kiowa Song: A Collaborative Ethnography. Tucson: University of Arizona
Press.

Lassiter, Luke E. 2005. “Collaborative Ethnography and Public Anthropology.” Current Anthropology 46 (1):
83–106.

Malinowski, Bronislaw. 1922. Argonauts of the Western Pacific: An Account of Native Enterprise and
Adventure in the Archipelagoes of Melanesian New Guinea. London: Routledge.

Millar, Kathleen M. 2018. Reclaiming the Discarded: Life and Labor on Rio’s Garbage Dump. Durham, NC: Duke
University Press.

Miller, Daniel. 2019. “Smartphones: The Cultural, Individual and Technical Processes That Make Them Smart.”
The Conversation. January 8, 2019. https://theconversation.com/smartphones-the-cultural-individual-and-
technical-processes-that-make-them-smart-106560.

Mintz, Sidney W. 1985. Sweetness and Power: The Place of Sugar in Modern History. New York: Viking.

Mohr, Sebastian. 2018. Being a Sperm Donor: Masculinity, Sexuality, and Biosociality in Denmark. New York:
Berghahn Books.

Morgan, Marcyliena. 2009. The Real Hiphop: Battling for Knowledge, Power, and Respect in the LA
Underground. Durham, NC: Duke University Press.

Access for free at openstax.org


1 • Bibliography 37

Nittle, Nadra Kareem. 2021. “Common Arab Stereotypes in TV and Film.” ThoughtCo. March 18, 2021.
https://www.thoughtco.com/tv-film-stereotypes-arabs-middle-easterners-2834648.

Orser, Charles E., Jr. 2007. The Archaeology of Race and Racialization in Historic America. Gainesville:
University Press of Florida.

Peregrine, Peter Neal. 2018. “Boas, Franz (1858–1942).” In The International Encyclopedia of Anthropology,
edited by Hilary Callan. Wiley Online Library. https://doi.org/10.1002/9781118924396.wbiea1299.

Ralph, Laurence. 2014. Renegade Dreams: Living through Injury in Gangland Chicago. Chicago: University of
Chicago Press.

Said, Edward. 1979. Orientalism. New York: Vintage Books.

Schaik, Carel van. 2004. Among Orangutans: Red Apes and the Rise of Human Culture. Cambridge, MA:
Belknap Press of Harvard University Press.

Singleton, Theresa A, ed. (1985) 2016. The Archaeology of Slavery and Plantation Life. New York: Routledge.

Stanley, Henry Morton. 1878. Through the Dark Continent; or, The Sources of the Nile around the Great Lakes
of Equatorial Africa and down the Livingstone River to the Atlantic Ocean. London.

Torres, Jada Benn, and Gabriel A. Torres Colón. 2021. Genetic Ancestry: Our Stories, Our Pasts. New York:
Routledge.

Tourigny, Eric. 2020. “Do All Dogs Go to Heaven? Tracking Human-Animal Relationships through the
Archaeological Survey of Pet Cemeteries.” Antiquity 94 (378): 1614–1629. https://doi.org/10.15184/
aqy.2020.191.
38 1 • Bibliography

Access for free at openstax.org


CHAPTER 2
Methods: Cultural and Archaeological

Figure 2.1 These archaeologists are working to uncover a fresco on a building in Pompeii, Italy. Pompeii was
famously covered in ash when nearby Mount Vesuvius erupted in 79 CE. The ash has preserved many structures and
artifacts from the time. (credit: “Pompeii Restoration Work” by Justin Ennis/flickr, CC BY 2.0)

CHAPTER OUTLINE
2.1 Archaeological Research Methods
2.2 Conservation and Naturalism
2.3 Ethnography and Ethnology
2.4 Participant Observation and Interviewing
2.5 Quantitative and Qualitative Analysis
2.6 Collections

INTRODUCTION Fieldwork is one of the most important practices of anthropology. While all of the subfields
of anthropology conduct fieldwork in some form to gather information, each subfield may use different
methods of conducting research. The concept of working in “the field” was traditionally based on the practice
of traveling to distant regions to study other cultures within their native environmental contexts. In recent
decades, “the field” has broadened to include diverse settings such as one’s hometown (as in urban
anthropology), the Internet (visual or virtual anthropology), or collections in university archives and museums
(ethnohistory or museum anthropology).
40 2 • Methods: Cultural and Archaeological

2.1 Archaeological Research Methods


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Describe archaeological techniques for uncovering artifacts.
• Explain the importance of context in making sense of artifacts and describe how researchers record content
while working in the field.
• Describe the law of superposition as used in the field of archaeology.
• Describe the different types of relative dating methods used by archaeologists.
• Identify and briefly define four absolute or chronometric dating methods.

Many people have an inherent fascination with the human past. Perhaps this fascination stems from the fact
that people recognize themselves in the objects left behind by those who have lived before. Relics of past
civilizations, in the form of human-made cultural artifacts, temples, and burial remains, are the means by
which we can begin to understand the thoughts and worldviews of ancient peoples. In the quest to understand
these ancient societies, human curiosity has sometimes led to fantastical myths about races of giant humans,
dragons, and even extraterrestrial beings. In the realm of archaeology, less speculative methods are used to
study the human past. Scientific approaches and techniques are the foundation of archaeology today.

Archaeological Techniques
In archaeology, the first step in conducting field research is to do a survey of an area that has the potential to
reveal surface artifacts or cultural debris. Surveys can be done by simply walking across a field, or they may
involve using various technologies, such as drones or Google Earth, to search for unusual topography and
potential structures that would be difficult to see from the ground. Cultural artifacts that are found may
become the basis for an archaeological excavation of the site. A random sampling of excavation units or test
pits can determine a site’s potential based on the quantity of cultural materials found. GPS coordinates are
often collected for each piece of cultural debris, along with notes on specific plants and animal found at the
site, which can be indicators of potential natural resources. Features such as trails, roads, and house pits are
documented and included in a full set of field notes. Government agencies have different protocols about what
constitutes an archaeological site; the standard in many areas is six cultural objects found in close proximity to
one another.

When preparing a site for excavation, archaeologists will divide the entire site into square sections using a grid
system, which involves roping off measured squares over the surface of the site. This grid system enables
archaeologists to document and map all artifacts and features as they are found in situ (in the original
location). All objects and features uncovered are assigned catalog or accession numbers, which are written on
labels and attached to the artifacts. These labels are especially important if artifacts are removed from the site.

Excavation is a slow process. Archaeologists work with trowels and even toothbrushes to carefully remove
earth from around fragile bone and other artifacts. Soil samples may be collected to conduct pollen studies.
Ecofacts—objects of natural origins, such as seeds, shells, or animal bones—found at a site may be examined
by other specialists, such as zooarchaeologists, who study animal remains, or archaeobotanists, who
specialize in the analysis of floral (plant) remains with an interest in the historical relationships between
plants and people over time.

Every cultural and natural object and feature is fully documented in the field notes, with its exact placement
and coordinates recorded on a map using the grid system as a guide. These coordinates represent an object’s
primary context. If uncovered objects are moved before documentation takes place, the archaeologist will lose
the archaeological context of that object and its associated data. Archaeological context is the key foundation
of archaeological principles and practice. In order to understand the significance and even age of artifacts,
features, and ecofacts, one needs to know their context and association with other objects as they were found
in situ. Objects that have been removed from their primary context are said to be in a secondary context.

Careful and proper documentation is vitally important. This information becomes part of the archaeological
record and guides and contributes to future research and analysis.

Access for free at openstax.org


2.1 • Archaeological Research Methods 41

FIGURE 2.2 This dig site in Vindolanda, England has yielded thousands of artifacts left behind by Roman occupiers
in the years 85 – 370 CE. (credit: “Digging Archaeology 4” by Son of Groucho/flickr, CC BY 2.0)

Archaeological Dating Methods


Establishing the age of cultural objects is an important element of archaeological research. Determining the
age of both a site and the artifacts found within is key to understanding how human cultures developed and
changed over time. Other areas of science, such as paleontology and geology, also use dating techniques to
understand animal and plant species in the ancient past and how the earth and animal species evolved over
time.

Relative Dating
The earliest dating methods utilized the principles of relative dating, developed in geology. Observing
exposed cliffsides in canyons, geologists noted layers of different types of stone that they called strata
(stratum in the singular). They hypothesized that the strata at the bottom were older than the strata higher up;
this became known as the law of superposition. According to the law of superposition, not just geological
layers but also the objects found within them can be assigned relative ages based on the assumption that
objects in deeper layers are older than objects in layers above. The application of the law of superposition to
archaeological fieldwork is sometimes called stratigraphic superposition. This method assumes that any
cultural or natural artifact that is found within a stratum, or that cuts across two or more strata in a cross-
cutting relationship, is younger than the stratum itself, as each layer would have taken a long time to form
and, unless disturbed, would have remained stable for a very long time. Examples of forces that might cause
disturbances in strata include natural forces such as volcanos or floods and the intervention of humans,
animals, or plants.

The law of superposition was first proposed in 1669 by the Danish scientist Nicolas Steno. Some of the first
applications of this law by scholars provided ages for megafauna (large animals, most commonly mammals)
and dinosaur bones based on their positions in the earth. It was determined that the mammalian megafauna
and the dinosaur bones had been deposited tens of thousands of years apart, with the dinosaur remains being
much older. These first indications of the true age of fossil remains suggested a revolutionary new
understanding of the scale of geological time.

It was eventually determined that if a specific set and sequence of strata is noted in several sites and over a
large enough area, it can be assumed that the ages will be the same for the same strata at different locations in
the area. This insight enabled geologists and archaeologists to use the structures of soils and rocks to date
phenomena noted throughout a region based on their relative positions. Archaeologists call this method
archaeological stratification, and they look for stratified layers of artifacts to determine human cultural
contexts. Stratigraphic layers found below cultural layers provide a basis for determining age, with layers
above assumed to be more recent than those below.
42 2 • Methods: Cultural and Archaeological

FIGURE 2.3 According to the principle of superposition objects found at deeper layers (called stratum) are older
than those found above. In this illustration, the pottery fragments in Stratum E can be assumed to be older than the
shell buttons found in Stratum C. The objects nearest the surface (aluminum can, plastic bottle) are obviously most
recent. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

Another method of dating utilized by archaeologists relies on typological sequences. This method compares
created objects to other objects of similar appearance with the goal of determining how they are related. This
method is employed by many subdisciplines of archaeology to understand the relationships between common
objects. For example, typological sequencing is often conducted on spearpoints created by Indigenous peoples
by comparing the types of points found at different locations and analyzing how they changed over time based
on their relative positions in an archaeological site.

Another form of typological sequencing involves the process of seriation. Seriation is a relative dating method
in which artifacts are placed in chronological order once they are determined to be of the same culture.
English Egyptologist, Flinders Petrie introduced seriation in the 19th century. He developed the method to
date burials he was uncovering that contained no evidence of their dates and could not be sequenced through
stratigraphy. To address the problem, he developed a system of dating layers based on pottery (see Figure
2.4).

Access for free at openstax.org


2.1 • Archaeological Research Methods 43

FIGURE 2.4 Petrie’s Egyptian pottery seriation method is built upon the observation that styles change with time.
Petrie arranged pottery artifacts into similar groups based on stylistic features and placed them along a relative
timeline based on these features. (credit: “Evolution of Egyptian prehistoric pottery styles, from Naqada I to Naqada
II and Naqada III” by W. M. Flinders Petrie and A. C. Mace/Wikimedia Commons, Public Domain)

Typological sequences of pottery, stone tools, and other objects that survive in archaeological sites are not only
used to provide dating estimates. They can also reveal much about changes in culture, social structure, and
worldviews over time. For example, there are significant changes in stratigraphy during the agricultural age, or
Neolithic period, at around 12,000 BCE. These changes include the appearance of tended soils, pollens that
indicate the cultivation of specific plants, evidence of more sedentary living patterns, and the increased use of
pottery as the storage of food and grain became increasingly important. Archaeological evidence also shows a
growing population and the development of a more complex cultural and economic system, which involved
ownership of cattle and land and the beginning of trade. Trade activities can be determined when pottery types
associated with one site appear in other nearby or distant locations. Recognizing the connections between
objects used in trade can shed light on possible economic and political interrelationships between neighboring
communities and settlements.

Chronometric Dating Methods


Chronometric dating methods, also known as absolute dating methods, are methods of dating that rely on
chemical or physical analysis of the properties of archaeological objects. Using chronometric methods,
archaeologists can date objects to a range that is more precise than can be achieved via relative dating
methods. Radiocarbon dating, which uses the radioactive isotope carbon-14 (14C), is the most common
method used to date organic materials. Once a living organism dies, the carbon within it begins to decay at a
known rate. The amount of the remaining residual carbon can be measured to determine, within a margin of
error of 50 years, when the organism died. The method is only valid for samples of organic tissue between 300
and 50,000 years old. To ensure accuracy, objects collected for testing are promptly sealed in nonporous
containers so that no atmospheric organic substances, such as dust, pollen, or bacteria, can impact the results.

Dating systems that measure the atomic decay of uranium or the decay of potassium into argon are used to
date nonorganic materials such as rocks. The rates of decay of radioactive materials are known and can be
measured. The radioactive decay clock begins when the elements are first created, and this decay can be
44 2 • Methods: Cultural and Archaeological

measured to determine when the objects were created and/or used in the past. Volcanic materials are
particularly useful for dating sites because volcanoes deposit lava and ash over wide areas, and all the material
from an eruption will have a similar chemical signature. Once the ash is dated, cultural materials can also be
dated based on their position relative to the ash deposit.

The technique of dendrochronology relies on measuring tree rings to determine the age of ancient structures
or dwellings that are made of wood. Tree rings develop annually and vary in width depending on the quantity
of nutrients and water available in a specific year. Cross dating is accomplished by matching patterns of wide
and narrow rings between core samples taken from similar trees in different locations. This information can
then be applied to date archaeological remains that contain wood, such as posts and beams.
Dendrochronology has been used at the Pueblo Bonita archaeological site in Chaco Canyon, New Mexico, to
help date house structures that were occupied by the Pueblo people between 800 and 1150 CE. The Laboratory
of Tree-Ring Research, based in Tucson, is the world’s oldest dendrochronology lab. Go on a tree-ring
expedition! (https://openstax.org/r/tree-ring)

The most effective approach for dating archaeological objects is to apply a variety of dating techniques, which
allows the archaeologist to triangulate or correlate data. Correlating multiple methods of dating provides
strong evidence for the specific time period of an archaeological site.

How It
Strategy What It Is How It Is Read Assumptions
Is Seen

Growth
Tree ring Count rings 1 ring = 1 year; no duplication or
Dendrochronology in life,
width pattern and measure missed rings; regional comparability
ring

Count beta
Radioactive Decay Half-life of 14C-12C decay known;
14C decay or 14C
decay and after exchange with atmosphere and
per unit
atom counting death productions rates constant
volume

TABLE 2.1 Chronometric Dating Techniques

2.2 Conservation and Naturalism


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Describe conservation efforts undertaken in the United States in the 19th century.
• Define salvage anthropology and describe its origins and methods.
• Provide an example of an anthropologist who used their research to help the people they were studying.
• Explain why museums can be said to have created exhibits reflecting limited interpretations and describe
efforts to correct this limitation.

Early Efforts
The conservation movement began in the 19th century as people in Europe and America began to realize that
human settlement and the exploitation of the world’s natural resources had led to the destruction or
endangerment of numerous animals, plants, and significant environments. Efforts began in the 1860s to
understand and protect the remaining natural landscapes and habitats. These efforts were partly motivated by
concern for wildlife and natural areas. However, also significant were the concerns of sporting organizations
and recreationists. The primary aim of early conservation efforts was to preserve significant natural
ecosystems for parks or wilderness areas so that sportspeople and outdoor enthusiasts would have places to
hunt, fish, and explore. Many areas preserved by these early efforts are still protected today, such as
Yellowstone and Yosemite National Parks in the United States.

Access for free at openstax.org


2.2 • Conservation and Naturalism 45

An element of this early period of conservation was the effort to collect specimens for display in natural history
museums. This collection effort was part of a movement known as naturalism, which seeks to understand the
world and the laws that govern it by direct observation of nature. The late 19th and early 20th centuries saw a
marked growth in naturalist collections worldwide as many cities and nations sought to establish and fill their
own natural history museums. These collections have been particularly useful to zooarchaeologists and
archaeobotanists, who use specimen collections of mammals, birds, fish, and plants to identify natural objects
and animal remains found at human burial sites. Many archaeology labs have collections of animal skeletons
for comparative anatomy, analysis, and identification (see Figure 2.5).

FIGURE 2.5 Collections of bones, such as this collection of specimens from various animal species housed at the
Wildlife Forensics Lab in Ashland, Oregon, serve as a useful resource for zooarchaeologists. (credit: “Wildlife
Forensics Lab” by US Fish and Wildlife Service Headquarters/flickr, Public Domain)

In addition to animal specimens, Native American baskets and other Indigenous art objects were collected and
placed in natural history museums. When visiting the Auckland Museum in Auckland, New Zealand, visitors
today encounter two large totem poles in the foyer. Northwest Coast totem poles are common in most older
museums throughout the world. These totem poles were gathered from America’s Northwest Coast in the late
19th and early 20th centuries as part of the worldwide conservation and naturalism movement. Most
museums sought to purchase such artifacts, but in some cases, artifacts were stolen when Indigenous owners
were unwilling to sell them. Many natural history museums also established dioramas depicting both
Indigenous peoples and animals in their “natural” world. The practice of installing dioramas of Indigenous
people is now heavily criticized because of the implication that Indigenous peoples are akin to animals and
plants. Many museums have stopped this practice and have even dropped the phrase natural history from
their names. However, the Smithsonian Institution’s National Museum of Natural History in Washington, DC,
and the American Museum of Natural History in New York both maintain the designation and still display
dioramas of Indigenous peoples.
46 2 • Methods: Cultural and Archaeological

FIGURE 2.6 This diorama of Native Americans is on display in the Indiana State Museum in Indianapolis, Indiana.
Such dioramas have come under criticism for the way that they depict Indigenous peoples and cultures. (credit:
“Native Americans – Indiana State Museum – DSC00394” by Daderot/Wikimedia Commons, Public Domain)

Salvage Anthropology
Connected to the collecting of Indigenous artifacts is a practice known as salvage anthropology. Salvage
anthropology was an effort to collect the material culture of Indigenous peoples in the United States and other
parts of the world who were believed to be going extinct in the later 19th century. During this period, many
anthropologists dedicated themselves to collecting material objects, stories, language lists, and ethnographies
from tribal peoples worldwide. Many collections were made through legitimate means, such as purchasing
objects or sitting down with collaborators (called informants in older anthropological vernacular) to record
traditional stories, but some collecting involved the theft of tribal cultural items or purchases from
intermediary traders.

Many of these anthropologists were hired by the Bureau of American Ethnology (BAE), a division of the
Smithsonian Institution, and spent considerable time living with Native peoples on the reservations that were
by then home to most Native Americans. Language was a special research focus for linguists and
anthropologists, as many Native languages were rapidly going extinct. Through analysis of language, an
anthropologist can understand the meaning of words and their context as well as gain a sense of a culture’s
philosophy and worldviews.

Anthropologists were not paid well to do this work for the BAE. Some began supplementing their income by
buying cultural objects at a low cost from the people they studied and selling those objects at a much higher
rate to museums. This practice is now acknowledged as unethical and exploitative. The anthropological
research of this period has also been criticized for focusing solely on cultural knowledge while ignoring the
hardships faced by the culture. For example, few anthropologists chose to help their subjects address the
circumstances of living in poverty on the reservations.

Leonard J. Frachtenberg was an anthropologist working during the salvage anthropology period who did take
action to help the people he was studying. Around the turn of the 20th century, Frachtenberg was conducting
research to collect the languages of the people living on the Siletz Reservation, in Lincoln County, on Oregon’s
coast. He worked extensively with collaborators from the Coos, Coquille, Lower Umpqua, and Alsea
tribes—some of whom were living at the Siletz Reservation and some who had returned to their native
lands—and published a series of oral histories based on his research. He also helped the tribes locate lost
unratified treaties from the 1850s and use those treaties to successfully sue the federal government. In the
treaties, the government had promised to pay the Indigenous peoples of Oregon’s coast for their ancestral land
if they peacefully relocated to the Siletz Reservation. The people upheld their part of the bargain, but they

Access for free at openstax.org


2.2 • Conservation and Naturalism 47

never received any payment. Frachtenberg helped a Coquille man named George Wasson travel to
Washington, DC, and locate copies of the treaties in the National Archives. In 1908, the tribes began the
process of successfully suing the federal government for payment for their lands. This process took some 40
years to complete for many tribes, and not all tribes have been fairly paid to this day.

Museum Collections
Most of the materials collected by anthropologists during the period of salvage anthropology ended up in
museums and university archives. Many natural history museums now display large dioramas featuring the
material objects of numerous tribes. Museum research libraries house extensive collections of manuscripts
and ethnographies. Archaeologists have contributed to these collections as well; many museums contain large
collections of human remains. Indigenous peoples have criticized these collections, especially the gathering of
human remains, which is seen as sacrilegious. Today, there are millions of sets of human remains (some full
skeletons, but most single bones) in museum repositories that have never been studied and perhaps never will
be.

Anthropologists spent so much of their time in the early period collecting that they had little time to study or
analyze what they found. Many collections were put in storage after the anthropologists who had gathered
them moved on to a new project or passed away. There are currently millions of material artifacts and
ethnographic manuscripts that have never been fully studied. These archived materials offer research
opportunities for anthropologists as well as for Indigenous peoples, who are making use of these collections to
help recover parts of their cultures that were lost due to the assimilation policies of the past 200 years.

One person who has taken advantage of these archives is linguistic anthropologist Henry Zenk. Zenk has spent
years studying the languages and cultures of the tribes of western Oregon, specifically the Chinook, Kalapuya,
and Molalla tribes. He conducted research with the Grand Ronde tribe in the 1970s and 1980s and became a
proficient speaker of Chinuk Wawa, a trade language spoken by tribes from southern Alaska to northern
California and as far east as Montana. He has taught the language at the Grand Ronde Reservation for nearly 30
years. He is also one of the experts on the Kalapuya languages, spoken by the Kalapuya tribes of the Willamette
and Umpqua Valleys, and in 2013, he began a project to translate the Melville Jacobs Kalapuya notebooks.

Melville Jacobs was an anthropologist from the University of Washington who studied the languages of the
Northwest Coast from 1928 until his death in 1971. He filled more than 100 field notebooks with information
on the languages of the peoples of western Oregon, with a special focus on Kalapuya. Jacobs published a book
of Kalapuya oral histories in 1945, Kalapuya Texts. He also worked with Kalapuya speaker John Hudson to
translate numerous texts prepared by earlier anthropologists Leonard Frachtenberg and Albert Gatschet.
Jacobs and Hudson were able to translate several of these previously gathered texts, but many remained
untranslated when Hudson died in 1953. Zenk, along with colleague Jedd Schrock, spent many years first
learning Kalapuya and then translating a set of the Jacobs notebooks that recorded the knowledge and history
of a Kalapuya man named Louis Kenoyer. In 2017, Zenk and Schrock published My Life, by Louis Kenoyer:
Reminiscences of a Grand Ronde Reservation Childhood. Zenk and Schrock’s work is a fine example of the
research possibilities offered by the existing work of previous anthropologists.

Zenk worked closely with the Grand Ronde tribe on this project and endeavored to make sure that the
translation of Kenoyer’s story would benefit the people of the tribe to help them to better understand their own
history. His research and work with members of the Grand Ronde tribe spanned 50 years, beginning with his
PhD project, which involved extensive work with Grand Ronde members, who at the time were not a federally
recognized tribe. In the 1990s, Zenk began working with the tribe to teach Chinuk Wawa to tribal members.
The tribe today has an extensive language immersion project to teach the language to young people. Zenk has
been a consistent influence, serving as advisor, teacher, master-apprentice instructor, and researcher. Zenk’s
work has helped the tribe recover parts of its culture and history that had been lost for many decades.
48 2 • Methods: Cultural and Archaeological

FIGURE 2.7 This page of an anthropologist’s field notebook from 1949 contains a travel itinerary for several
months. Contemporary anthropologists are likely to have such information in digital format. (credit: “Field Notes –
Mexico, 1949 (Page 180) BHL46264382” by James Arthur Peters/Biodiversity Heritage Library/Wikimedia
Commons)

PROFILES IN ANTHROPOLOGY

Albert Gatschet
1832–1907

FIGURE 2.8 Albert Gatschet was a Swiss-American ethnologist who pioneered the scientific study of Native

Access for free at openstax.org


2.2 • Conservation and Naturalism 49

American languages. Here he is at age 61. (credit: “PSM V41 D306 Albert S Gatschet” by Popular Science
Monthly/Wikimedia Commons, Public Domain)

Personal History: Albert Gatschet was a Swiss philologist and ethnologist who emigrated to the United States
in 1868. He had a great interest in linguistics and Native American languages, and he gained attention in 1872
for his comparative analysis of 16 southeastern tribal vocabularies, which opened up new areas of research in
linguistics. In 1877, he was hired to work on the Geographical and Geological Survey of the Rocky Mountain
Region as an ethnologist. He also collected many notebooks of languages from Native peoples in California and
Oregon. He is most noted for his studies of the languages of the southeastern tribes and his ethnography of the
Klamath Tribes of Oregon.

Gatschet was fluent in numerous languages and published in English, French, and German in the United
States and Europe during his career. He also became quite fluent in numerous Native languages. His first large
work was Orts-etymologische Forschungen aus der Schweiz (Etymological research on place names from
Switzerland, 1865–1867), a study of Swiss place names that is still the standard authority today.

Area of Anthropology: Philology, ethnology, linguistics

Accomplishments in the Field: One of Gatschet’s most significant analyses was of the southeastern tribal
languages, principally the Timucua language of northern Florida. Based on analysis of the notes of the Catholic
priest Father Pareja, who had collected language texts from the Timucua people in 1612–1614, Gatschet
determined that Timucua was a distinct language group that had gone extinct. Gatschet also examined the
Catawba language of South Carolina, concluding that it was related to the Siouan languages of the western
Great Plains. From 1881 to 1885, Gatschet worked in Louisiana, discovering two new languages and
completing ethnographic descriptions of the southern tribes. In 1886, he found the last speakers of the Biloxi
and Tunica languages and related them to the Siouan languages as well. He published his studies of the Gulf
tribes in the two-volume work A Migration Legend of the Creek Indians (1884, 1888).

In 1877 and 1878, Gatschet spent time among the tribes of the Grand Ronde Reservation in Oregon. He
collected some of the first professional field notes on the Kalapuya, Molala, and Shasta languages from some of
the last speakers, and he published and made notes about the Kalapuya mounds. Upon leaving the reservation,
he spent time researching the traditions of the Tualatin Kalapuya people in their traditional lands in the
Tualatin Valley. He then went to the Klamath Reservation, where he collected field notes on the Klamath
language. He worked his field notes into a two-part work, The Klamath Indians of Southwestern Oregon (1890),
volume 2 of the US Department of the Interior’s Contributions to North American Ethnology.

Gatschet was commissioned by the Bureau of American Ethnology (BAE) in 1891 to investigate the Algonquian
people of the United States and Canada, a study he never fully completed. Illness forced him to retire, but near
his death, he remained engaged in studies of Chinese languages.

After his death, his wife, Louise Horner Gatschet, sold his field notes to the BAE. She was also hired by the BAE
to help translate much of his work. Gatschet’s letters mention his wife being with him throughout his travels;
she likely contributed in numerous ways to his field studies.

Importance of His Work : Gatschet was one of the first professional anthropologists to visit many tribes and
was able to collect ethnographies and narratives from peoples who were gone within the next decade. He
analyzed language families in the field and provided early frameworks of connected languages. Gatschet’s
work is fundamental to the study of the languages of western Oregon and the southeast Gulf area of the United
States. His professional work, which applied rigorous methods to collect Native languages, predates much of
the work of Franz Boas, who is credited with implementing scientific methods in the study of human societies.

Interpretation and Voice


There is increasing acknowledgement of the role of interpretation in the study of the human past. Although
ideally grounded in well-conducted research and the best evidence available at the time, all conclusions about
what might have been are based on the interpretations proposed by the authors of history. The backgrounds
and viewpoints of those conducting research and publicizing findings play a significant role in the conclusions
50 2 • Methods: Cultural and Archaeological

they reach and share with other scholars. Interpretation and perspective are affected by many factors,
including racial category, nationality, religious beliefs, social status, political affiliation, ambitions, and
education. For many years, anthropological studies were almost always conducted by White, male scholars
who grew up in the Northern Hemisphere and were educated in the same system. These common
backgrounds represent a significant interpretive bias.

After being accessioned into museums, many collections of cultural artifacts have not been altered in more
than 100 years. When these material objects were initially placed on display, choices about their arrangement
and the written descriptions that accompanied them were made by museum curators. Most of these curators
did not reach out to the originators of the artifacts or their descendants for input, and many exhibits do not
accurately depict or describe the objects on display. Museum exhibits have been found to contain inaccurate
information about objects’ material composition, makers, tribal cultures, collection sites, and proper use.
Many other display objects are lacking this information altogether.

Several museums are now seeking the help of Native people to better understand and more accurately tell the
story of their collections. These Native perspectives are correcting misconceptions about the meaning and
context of cultural artifacts and providing correct information about basic things such as the materials and
processes used in the objects’ production. Native input is also guiding museums in making choices about how
objects are arranged and displayed. This input has been invaluable in helping museums more accurately tell
the stories and display the context of the peoples who originally created the objects on display.

2.3 Ethnography and Ethnology


LEARNING OUTCOMES

By the end of this chapter, you will be able to:


• Identify early anthropological practices pertaining to ethnography.
• Define ethnology and provide examples of how it is used in anthropology.
• Describe efforts to achieve multiple perspectives in anthropological research.
• Define feminist anthropology and describe its aims.

The Development of Ethnography and Ethnology


As discussed in What is Anthropology? ethnography is a method used by cultural anthropologists to create a
description of a culture or society. Ethnographers gather and utilize information from many sources, such as
fieldwork, museum collections, government records, and archaeological data. In the 19th century, a form of
ethnography developed that was called armchair anthropology, in which theories about human societies and
human behaviors were proposed solely based on secondhand information. Lewis Henry Morgan is a well-
known practitioner of this type of research. The content of his most famous publication, League of the Ho-dé-
no-sau-nee, or Iroquois (1851), was gathered primarily from other books he read. Morgan did meet with Native
peoples at various times in his career, but he did not conduct ethnographic research among the Iroquois
before writing League of the Ho-dé-no-sau-nee, or Iroquois.

In the later 19th century, numerous anthropologists and other scholars undertook research projects with
hundreds of tribes throughout the Americas, many of them by then living solely on federal reservations. Many
of these researchers were influenced by Columbia University professor Franz Boas, a German scientist who
was originally trained as a physicist but became most famous as an anthropologist. Boas insisted that scholars
obtain ethnographical information directly from the peoples they aimed to write about, rather than collecting
information from other published sources. Boas quickly established himself as a leader in the field of
anthropology and eventually took an associate role at the federal Bureau of American Ethnology.

Access for free at openstax.org


2.3 • Ethnography and Ethnology 51

FIGURE 2.9 Franz Boas is credited with establishing the standards of field research that became the foundation of
contemporary anthropological practices. Here he is in 1915, 57 years old. (credit: “Franz Boas” by Canadian
Museum of History/Wikimedia Commons, Public Domain)

Boas advocated for and published in all four fields of anthropology and asked many key questions in his
scholarship. In his 1907 essay “Anthropology,” Boas identified two basic questions for anthropologists: “Why
are the tribes and nations of the world different, and how have the present differences developed?” (Boas
[1974] 1982, 269). Boas was responsible for hiring scholars and sending them out into the field to collect
information about various Indigenous peoples. His standards of field research became the foundation of the
contemporary science of anthropology.

One area of interest for early anthropologists was the similarities and differences between various Indigenous
societies. This interest in comparison led to a branch of anthropology called ethnology, which is a cross-
cultural comparison of different groups. In early anthropology, ethnology’s aim was to understand how various
Indigenous societies were related to one another. This included the relations among language dialects, dress,
and appearance and to what degree and in what direction various tribes had migrated from one location to
another. Early anthropologists explored these questions with the hope of tracking changes in tribal cultures.
Another leading concern was how Native peoples initially got to the Americas. Anthropologists have used the
practices of ethnology to establish relationships and shared cultural elements that help illuminate migration
patterns of peoples from the “old” to the “new” world. Ethnology is still a common practice in linguistics,
archaeology, and biological anthropology.

Some additional uses of ethnology are fused with archaeological methods and analysis. Ethnoarchaeology is a
form of archaeology in which, following methods largely created by American archaeologist Lewis Binford,
archaeologists access ethnographic information about recent or existing human cultures to draw conclusions
about human cultures in the archaeological past. In Binford’s 1978 study Nunamiut Ethnoarchaeology, he
draws comparisons between the ways in which contemporary Indigenous peoples disposed of animal remains
and the evidence observed in Nunamiut refuse sites. These comparisons inform a model that is used to
understand more about how Indigenous peoples’ ancestors may have disposed of remains in the past. Such
models are not perfect, but many Indigenous cultures have maintained aspects of their culture to the present
day.
52 2 • Methods: Cultural and Archaeological

Perspective and Interpretation in Ethnography


Ethnography is still commonly used by cultural anthropologists. Practitioners today consult multiple
informants during their research in order to gather a variety of perspectives on a culture or society. No one
person has a full or authoritative view of their own culture; multiple viewpoints are essential to a full
description. Many early anthropological studies only invited male perspectives, introducing a male bias into
the resulting ethnographies. Now, anthropologists deliberately seek varied perspectives, consulting people of
different genders and ages and who occupy different roles.

Anthropologists can introduce significant bias into an ethnography. The most challenging aspect of fieldwork
in cultural anthropology is to observe and study another culture without bias. Having an ethnocentric or etic
perspective means someone is judging a culture according to the standards of their own culture and belief
system. To observe a culture from the perspective of the people being researched is to have an emic
perspective. For anthropologists to be effective researchers, they must be able to observe and gather data
from unbiased and emic perspectives. In addition, an anthropologist’s interpretation of the information
gathered can significantly alter their research findings. Earlier anthropologists were primarily male and White,
so their findings were based on interpretations made through these lenses. Feminist anthropology attempts
to address this male bias. Feminist anthropology is recognized as having begun as early as the 1850s, with
attempts made (by male anthropologist) to include more information on women in their ethnographic
research. In the 1920s, female anthropologists such as Zora Neale Hurston and Ruth Benedict began
publishing in the field, but not until the 1928 publication of Margaret Mead’s Coming of Age in Samoa did a
female anthropologist gain prominence.

FIGURE 2.10 This U.S. postage stamp honors anthropologist Margaret Mead. Mead was one of the first female
anthropologists to be acknowledged for her work and insights. (credit: “Margaret Mead Stamp” by John Curran/
flickr, CC BY 2.0)

Women’s contributions and perspectives became much more pronounced in the later parts of the 20th
century. Feminist anthropologists seek not only to claim a role for themselves in the field equal to that offered
to men but also to expand the focal points of anthropological inquiry to include areas of life such as family,
marriage, and child-rearing, as well as the economic and social roles played by women. The dominance of
male anthropologists had biased analysis of human societies toward male-dominated roles and activities.

Access for free at openstax.org


2.4 • Participant Observation and Interviewing 53

Many early archaeological research, for example, assigned no role to women in early societies or assumed that
women’s roles were limited to maintaining households and raising children. Evidence of women’s subsistence
and economic activities was either not looked for or ignored. It was also assumed that women in early societies
had subservient roles to men, when in fact most early societies have now been found to be very egalitarian,
with equal status accorded to women and men. Feminist anthropology has both expanded research to include
women’s roles and aimed to understand the gender roles in other societies on their own terms, rather than
according to the gender roles of the researcher’s own society.

Other perspectives emerged in anthropology in the 1970s as more members of minority groups began
entering the field. One category of minority voices that has been a significant asset to anthropology is that of
people with Indigenous ancestors. Practitioners with this type of background are part of a subfield called
Indigenous anthropology. Indigenous anthropology is discussed in detail in Indigenous Anthropology.

2.4 Participant Observation and Interviewing


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define participant observation and identify best practices associated with it.
• Describe what makes a good informant for anthropological research.
• Describe best practices for conducting an interview from an unbiased and emic perspective.
• Explain the concept of ownership of cultural information.
• Identify the rights of study informants.
• List practices required by institutional review boards before research can begin.
• Describe the aim of long-term research projects in anthropology.

Participant Observation
Working in the field often places anthropologists in settings very different from what they are familiar with.
Upon first arriving at an unfamiliar field location, it is common for anthropologists to feel out of place and
uncomfortable as they adjust to a new culture and environment. Many anthropologists keep a daily log of their
feeling and impressions in their new environment. Researchers studying other cultures practice a method
called participant observation, which entails directly participating in the activities and events of a host
culture and keeping records of observations about these activities.

Researchers may create various types of records of their interactions as participants and their observations
about the host culture and environment. These might take the form of field notebooks, computer files, digital
recordings, photographs, or film. Researchers working in the field may also collect objects that will remind
them of the culture they are studying, often memorabilia such as maps, tourism brochures, books, or crafts
made by the people they are observing.

Some researchers regularly record impressions of activities while they are occurring so that they do not forget
to make note of important aspects of the culture. But many researchers will wait to take photos, draw images,
or write in their notebooks until after an activity is over so that they do not disturb the culture through their
efforts at documentation. In either case, it is important that researchers be respectful and responsible and
always ask for permission from subjects before taking photos or recordings. Many researchers will have
gathered signed permission from their subjects before beginning their research and will work with a
documented plan that has been approved by their institution before going into the field.

Interviewing Informants
An important source of information about a culture is interviews with various people who grew up in that
culture. Interviews can be uncomfortable for people, and it is important that researchers do all they can to help
subjects feel at ease. Researchers will normally conduct an interview in a familiar space for the informant,
such as the informant’s home. They will help the subject ease into the interview by participating in
introductory and hosting protocols followed in that culture when a visitor comes to someone’s home. The
researcher will start off the interview with the exchange of pleasant comments and will introduce themselves
54 2 • Methods: Cultural and Archaeological

by explaining who they are, where they come from, and why they are doing this research. Then the interview
may commence.

Interviews can be short or long, and there may be follow-up meetings and further interviews based on how
knowledgeable the informant is. Many informants are chosen because they are deeply conscious of multiple
aspects of their culture. This type of insider information is vitally important to an anthropological research
project. In addition to interview questions, survey questions may also be asked during these meetings. The use
of recording equipment, for both audio and video recordings, is common during interviews. However, such
equipment may be considered intrusive by some, and their use is always at the discretion of the informant.
Express permissions must always be obtained both to create a recording and to use a recording in future
projects.

FIGURE 2.11 Ethnographic researchers engage with the cultures they are studying by spending time with their
people. Here Josphat Mako, a Maasai man, greets Stuart Butler. Butler spent two months with Mako, walking
between Maasai villages and visiting with residents to learn about both traditional customs and contemporary
practices and challenges. (credit: “2015 06 24 Walking with the Maasai JPEG RESIZED 0025” by Make It Kenya
Photo/Stuart Price/flickr, Public Domain)

Ethical Considerations
Contemporary sociocultural researchers and anthropologists must follow protocols established by an
institutional review board (IRB) as well as any research protocols specific to the culture being researched. For
social science research, IRBs are committees housed within a university that must review and approve
research plans before any research begins. There may also be a parallel review process within the host culture.
The proposed research is normally fully planned out before the review process can begin, with specific
information about the type of research that will be conducted, including examples of questions to be asked,
potential risk factors to subjects, plans for emotional support for subjects, means of protecting the identity of
subjects, language used to fully disclose the intent of the project to subjects, and the final plan for archiving the
research data. Many Indigenous nations have their own research protocols, and foreign countries will have
their own research protocols and processes for securing permission to conduct research as well.

Researchers conducting sociocultural, medical, or clinical studies must gain written consent for all interviews
from their informants, and they must be transparent as to why they are conducting research and how it will be
used in the future. There are normally various levels of protocols pertaining to research, based on the potential
to cause stress or harm to the subjects. At the highest level, full disclosure and signed permission as well as
complete anonymity of the subjects involved in the project are required. A research plan should also specify
whether recordings, notes, and data will be archived for future use or destroyed at the end of the project.
Content gathered from research may make its way into articles or books or become part of a vast body of
anonymous data available to other researchers. These possibilities should be discussed with collaborators.
Collaborators are usually anonymous unless they choose to allow their names to be used. Many researchers
now assign to their subject culture significant rights to review reports and edit and correct erroneous

Access for free at openstax.org


2.5 • Quantitative and Qualitative Analysis 55

information and interpretations as well as ownership rights of the final product and the research data.
Alternately, researchers may destroy research data once the project is over so that it cannot be used in ways
other than what was originally intended.

Long-term research projects are becoming the norm for many professional researchers, who establish trusting
relationships with collaborators over the length of their careers. During the early years of anthropology, it was
almost unheard of for researchers to establish long-term relationships with the subjects of their research, but
many scholars began to view short-term relationships as exploitative. Long-term relationships involve a
regular return to the subject culture, on an annual or semiannual basis, to follow up on projects and programs.
Researchers often include their subjects in the planning and administration of their projects and will at times
seek a research objective based on the needs of their subjects. This type of research is more open-ended and
often has an applied focus, seeking to solve problems and issues identified as significant by the collaborating
culture. Those who engage in this type of research make it a primary aim to help the collaborating culture
rather than to seek information pertinent to their personal projects.

This type of open-ended research has been developed in response to the criticisms of Indigenous scholars
such as Vine Deloria Jr., who questioned whether early anthropologists did anything beneficial for the people
they studied. A researcher working in this fashion will listen closely to the concerns expressed by those they
are studying and aim to identify a project that will ultimately help the collaborating culture address issues
identified as important, either by directly working toward a solution or by offering significant insights into the
causes and subtleties of the issue. The researcher will include members of the culture in their team, and the
results of the research will be given to the people for their use. Researchers working in this manner may still
publish their findings, but the subject community will be part of the decision-making regarding what is
important and what should and should not be published. The subject community will also have control over
any projects that develop based on the findings. In some cases, the researcher is required to submit all
manuscripts intended for publication to a committee formed by the collaborating culture for review,
correction, and approval. Many Indigenous anthropologists who are tribal members are required to submit
their publications to their tribal council for approval before they publish.

Contemporary anthropological researchers often assign ultimate ownership of the material they collect to the
culture-bearers who provided the information. In fact, there are scholars today who, when publishing findings,
assign authorship to the community they worked with and assign themselves the role of editor or compiler. An
example is the text Chinuk Wawa: Kakwa nsayka ulman-tilixam laska munk-kemteks nsayka / As Our Elders
Teach Us to Speak It, which is authored by the Chinuk Wawa Dictionary Project and published by the
Confederated Tribes of the Grand Ronde Community of Oregon, with the scholar Henry Zenk acknowledged as
the compiler of the information. Intellectual property protocols in many countries now assume that ownership
of ethnographic content is assigned to the informants. Informants have rights, both legally and per IRB
policies, to both participate and not participate in a study and to have their data removed from a study if they
choose. Ethical researchers will listen to their informants, and if they are at all worried about the effect their
findings will have on their informants or other people, they will either pull the data out of the study or find a
way to make it completely anonymous. No researcher wants to have their informants adversely affected by
their involvement in a research project. The IRB-informed consent paperwork, which must be signed by all
informants, should address these concerns and allow the informants to freely choose their level of
participation.

2.5 Quantitative and Qualitative Analysis


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Identify differences between quantitative and qualitative information.
• Provide an example of how an anthropologist might model research findings.
• Describe the steps of the scientific method.

Differences between Quantitative and Qualitative Information


Quantitative information is measurable or countable data that can provide insight into research questions.
56 2 • Methods: Cultural and Archaeological

Quantitative information is one of the most direct ways to understand limited, specific questions, such as how
often people in a culture perform a certain action or how many times an art form or motif appears in a cultural
artifact. Statistics created from quantitative data help researchers understand trends and changes over time.
Counts of cultural remains, such as the number and distribution of animal remains found at a campsite, can
show how much the campsite was used and what type of animal was being hunted. Statistical comparisons
may be made of several different sites that Indigenous peoples used to process food in order to determine the
primary purpose of each site.

In cultural research, qualitative data allows anthropologists to understand culture based on more subjective
analyses of language, behavior, ritual, symbolism, and interrelationships of people. Qualitative data has the
potential for more in-depth responses via open-ended questions, which can be coded and categorized in order
to better identify common themes. Qualitative analysis is less about frequency and the number of things and
more about a researcher’s subjective insights and understandings. Anthropology and other fields in the social
sciences frequently integrate both types of data by using mixed methods. Through the triangulation of data,
anthropologists can use both objective and frequency data (for example, survey results) and subjective data
(such as observations) to provide a more holistic understanding.

Modeling
Many anthropologists create models to help others visualize and understand their research findings. Models
help people understand the relationships between various points of data and can include qualitative elements
as well. One very familiar model is a map. Maps are constructed from many thousands of data points projected
onto a flat surface to help people understand distances and relationships. Maps are typically two-dimensional,
but we are of course all familiar with the three-dimensional version of a world map known as a globe. Maps
and globes are built on data points, but they also include qualitative information, such as the colors used to
represent various features and the human-assigned names of various geographical features. Other familiar
types of models include graphs, calendars, timelines, and charts. GPS is also a significant modeling tool today.

GPS, or the Global Position System, is increasingly used in archaeology. A model of a research site can be
created using computer programs and a series of GPS coordinates. Any artifacts found or important features
identified within the site can mapped to their exact locations within this model. This type of mapping is
incredibly helpful if further work is warranted, making it possible for the researcher to return to the exact site
where the original artifacts were found. These types of models also provide construction companies with an
understanding of where the most sensitive cultural sites are located so that they may avoid destroying them.
Government agencies and tribal governments are now constructing GPS maps of important cultural sites that
include a variety of layers. Layering types of data within a landscape allows researchers to easily sort the
available data and focus on what is most relevant to a particular question or task.

Wild food plants, water sources, roads and trails, and even individual trees can be documented and mapped
with precision. Archaeologists can create complex layered maps of traditional Native landscapes, with original
habitations, trails, and resource locations marked. GPS has significant applications in the re-creation of
historic periods. By comparing the placements of buildings at various points in the past, GPS models can be
created showing how neighborhoods or even whole cities have changed over time. In addition, layers can be
created that contain cultural and historic information. These types of models are an important part of efforts
to preserve remaining cultural and historic sites and features.

The Science of Anthropology


Anthropology is a science, and as such, anthropologists follow the scientific method. First, an anthropologist
forms a research question based on some phenomenon they have encountered. They then construct a
testable hypothesis based on their question. To test their hypothesis, they gather data and information.
Information can come from one or many sources and can be either quantitative or qualitative in nature. Part of
the evaluation might include statistical analyses of the data. The anthropologist then draws a conclusion.
Conclusions are rarely 100 percent positive or 100 percent negative; generally, the results are somewhere on a
continuum. Most conclusions to the positive will be stated as “likely” to be true. Scholars may also develop
methods of testing and retesting their conclusions to make sure that what they think is true is proven true
through various means. When a hypothesis is rigorously tested and the results conform with empirical

Access for free at openstax.org


2.6 • Collections 57

observations of the world, then a theory is considered “likely to be accurate.” Hypotheses are always subject to
being disproven or modified as more information is collected.

2.6 Collections
LEARNING OUTCOMES

By the end of this section, you will be able to:


• Identify and explain the issues and needs of archival collections.
• Identify and explain the issues and needs of three-dimensional collections.
• Describe current controversies regarding ownership of anthropological artifacts and human remains.
• Recall two pieces of legislature pertaining to questions of ownership.
• Define provenance and describe its importance in anthropology.

Not all anthropological research is done in the field. There is much to be learned from the collections of
manuscripts and artifacts housed in universities and museums. These collections make it possible for
anthropologists to study human cultures within the setting of special research laboratories that have been
designed to preserve and organize materials collected and perhaps interpreted by scholars of the past.

Archives
Archival collections contain published, re-created, or original manuscripts that are deemed significant enough
to be placed in conditions designed to preserve them against damage or loss. Such collections may contain
correspondence, maps, drawings, original drafts of books, rare books, or other papers and media that need
special care. Photographs are a major resource in many archives, and they need special handling. Preservation
policies of archival collections include practices such as keeping resources out of direct sunlight and away
from moisture.

While archives offer researchers a great range of valuable resources, they typically impose rather strict
policies on those wishing to access these resources. Researchers typically must wear gloves when handling
materials to prevent damage from the oils and acidity of human skin. Normally, archival collections do not
circulate (i.e., cannot be removed from the host site), and researchers may have to apply for permission to
enter the site or use any information. Archives may charge varying rates to make copies of material or to use
images of the resources in their collection for publication. To access some archives, researchers must plan
ahead by scheduling a time to visit and making previous arrangements to access specific collections. Some
sites do not allow researchers to scan materials using flatbed scanners, instead stipulating the use of non-flash
photography or overhead scanning. Some archives do not allow the patron to scan, photograph, or copy a
manuscript in any way, with all arrangements for copies and reproductions having to go through the archive’s
staff.

The first step in archival research is typically to review a list or similar finding aid that indexes and describes
the resources available in a collection. These descriptive aids can help researchers determine whether a
collection contains resources that fit their needs and can make a visit to a selected archive more efficient and
worthwhile. Finding aids have become so well constructed that they may provide researchers with enough
information to enable the researcher to request copies of specific materials and avoid the effort and expense of
traveling to the archive in person. Most archives offer downloadable finding aids of their most important
collections on their websites, and there may be additional printed finding aids available on request. Most
archives will make requested copies for a moderate fee and will mail or email researchers a packet of the
reproduced materials. The cost of procuring such copies is almost always much less than the cost of traveling
to an archive site and paying for housing and meals. However, if a collection is potentially full of material
important to a research project, it may be better to visit in person.

Three-Dimensional Collections
Three-dimensional collections of objects such as basketry and pottery are normally housed separately from
manuscript collections. Such collections may host tens of thousands of individual cultural objects. These
collections typically require much more care and management than manuscript materials. Extensive planning
goes into determining the best way to contain and store each type of object in order to slow deterioration over
58 2 • Methods: Cultural and Archaeological

time, with special attention paid to both the temperature and the moisture levels in storage areas. Handwoven
baskets will be supported so that their fibers are not under stress, and all organic objects will have been
previously frozen, perhaps several times, to destroy any insects that may live in the fibers. Collections of
animal and human remains utilized by biological anthropologists or archaeologists must be properly stored
and controlled against further degradation by reducing temperatures and maintaining moisture controls.
Some very ancient organic collections may need to be chemically stabilized so they do not degrade. Objects
made from organic materials—such as wooden canoes, basketry, reed sandals, or human remains—are
particularly prone to degradation. Organic artifacts that have been sealed away from contact with the air for
centuries, such as boats found on the bottom of a river or lake, will degrade fast once exposed to the air, so they
may be kept permanently frozen or preserved with an ammonium glycol solution to stabilize decay.

FIGURE 2.12 This pair of yucca sandals, collected in 1875, is an example of an organic artifact of the Southern
Paiute people. Yucca is a perennial plant with large tough leaves that can be used for various purposes. (credit:
“Sandals, Southern Paiute, yucca, collected in 1875 – Native American collection –Peabody Museum, Harvard
University – DSC05570” by Daderot/Wikimedia Commons, CC0)

All objects in collections storage must be well organized to make them accessible for further research
opportunities. Collection materials that have been used to make claims about human experience or evolution
must remain accessible to future researchers in case there are challenges or additional questions about their
findings. In addition, if an anthropologist who donated and is responsible for overseeing a collection at one
institution should die or move to another research institute, there needs to be a plan for the period of retention
for the collection, or the time that the collection will remain in the archive. Many biological and cultural
collections have been preserved in repositories since the day they were collected, with no plans to ever remove
them from an archive. There are collections in the Smithsonian Institution that have been there since the
institution was built in the 1850s. These collections continue to grow at museums and universities around the
world.

In the early 20th century, many museums adopted the practices of painting objects with lacquer and spraying
organic collections with pesticides such as DDT to prevent insect damage. These solutions were proven to
ultimately be harmful. Lacquer tends to alter the color and chemical structure of objects and is thus not a good
preservation material, and DDT and other pesticides pose health threats to humans. Both museum staff and
tribal members who receive repatriated objects and human remains are very concerned about the hazards
these chemicals pose to humans—and to the environment, if they should be reburied. Efforts to clean many
collections are underway.

Ownership
A question being asked by both anthropologists and subjects of research today is who owns the objects housed
in material collections. In the past, anthropologists or their host institutions assumed ownership of anything

Access for free at openstax.org


2.6 • Collections 59

they collected, along with the right to publish images of materials and sign over ownership of the objects to
collections repositories. In recent decades, tribal peoples and other subjects of research have begun asking
questions about whether such objects really should be considered the property of these repositories. Many of
these artifacts were not even collected by scientists but rather donated or sold by collectors, some of whom
removed the artifacts from burial sites. Artifact hunting is a common cultural practice in some countries, such
as Peru, where many people dig in Inca sites to locate artifacts to sell.

Questions of ownership become particularly pressing when the objects in question are human remains. Until
the 1960s, tribal peoples in the United States had little or no power to repatriate their ancestors. Repatriation
is the process of restoring human remains and/or objects of religious or cultural importance to the peoples
from whom they originated. In the United States, repatriation is executed under the Native American Graves
Protection and Repatriation Act (NAGPRA), passed into law in 1990. Prior to 1990, Indigenous peoples in the
United States had no legal means to claim return of any of the millions of human remains that had been
collected and placed in museums and archaeological collections since the 19th century.

Another important piece of legislature is the National Historic Preservation Act (NHPA), passed in 1966. The
act was passed to ensure that federal agencies would identify and take actions to protect and preserve the
nation’s historic sites and locations. It especially impacted Indigenous communities and their cultural and
historical resources. Section 106 of the NHPA requires that federal agencies follow a formal review process
before undertaking any type of development project (36 CFR 800). This process includes identifying what the
actual undertaking is, such as the development of a road or other major capital project. Once this is
established, the agency must make a good-faith effort to identify any historic resources (50+ years of age) in
the area and determine if they are eligible for protection under the NHPA. After this identification measure is
completed, the agency must initiate consultation with the state historic preservation officer (SHPO) or tribal
historic preservation officer (THPO) and other interested groups and individuals. This step can include a
variety of meetings or activities and a period of notification that a project is going to commence, during which
feedback is requested by the lead federal agency. Public meetings might be held, with speakers selected to
introduce and describe the project. During the consultation period, correspondence and feedback is welcomed
from concerned tribes, institutions, or individuals. Tribes and other community groups with an interest in any
cultural objects likely to be found on the site are required to be consulted. Successful consultation often takes
place during the earliest planning stages of a project. Lack of early consultation can lead to a failure to identify
historic resources of cultural and religious importance.

The process places the burden of determining the potential effects of the project on the federal agency,
according to three established categories: no potential to effect, no adverse effect, and adverse effect. The
agency must then seek concurrence from appropriate SHPOs and THPOs and potentially other consulting
parties. If there is an adverse effect, the agency, the SHPO and/or THPO, and other consulting parties will
negotiate mitigation terms and solidify them into a memorandum of agreement to ensure completion of the
agreed-upon mitigation measures. In most cases, Native groups do not believe that archaeological excavations
alone are an appropriate mitigation measure, but each community has its own interpretation of what is
appropriate.

Generally, anytime a road is built or a building is constructed, there needs to be a section 106 review of the
project because of the likelihood of encountering Native American cultural sites in almost all locations in the
United States. Through the consultation process and cooperation between SHPOs and THPOs, decisions are
made as to the status and disposition of any cultural objects recovered from cultural sites. Tribes typically
advocate for the non-disturbance of human remains and the return of cultural objects to the concerned tribes.
The NHPA is not perfect, as it does not completely halt construction that will destroy a cultural site and does
not apply to collections placed in repositories before 1966.

In the early 20th century, the United States made it illegal for nonscientists to remove artifacts from
archaeological sites on federal lands under a law called the American Antiquities Act (1906). More recently,
NAGPRA made it possible for tribes to repatriate objects covered under the act, such as human remains and
funerary objects. Under this law, more than 20,000 sets of remains had been repatriated as of 2010, but
millions of artifacts and sets of additional remains are still in repositories. In addition, there are human
remains and funerary objects of US origin in collections worldwide that are not subject to NAGPRA
60 2 • Methods: Cultural and Archaeological

repatriation.

One problem surrounding repatriation is that many artifacts and remains lack clear provenance, or detailed
information about where they were found. Lack of clear provenance also limits an object’s usefulness to
researchers. In many cases, wide regions are provided as the origin of an artifact, making it unclear which
specific tribal culture it relates to. Objects that, for example, are labeled as coming from “New York” may have
been created by members of dozens of tribes or bands of tribes. In general, the more specific a provenance is,
the better. Narrowing an object down to Buffalo, New York, reduces its possible tribal sources to just a few.
Objects that have too broad of a context are nearly impossible to repatriate because repatriation is supposed to
return an object or human remains to the original tribe. In 2010, NAGPRA was expanded to allow for groups of
tribes to repatriate objects of wide regional association back to a previously agreed-upon reburial or
repatriation location. Under this expanded version of the law, a greater number of objects and human remains
will be able to be returned to their communities.

Concerns about ownership have also been raised regarding the ethnological and ethnographic research
collected in millions of documents in hundreds of research collections around the world. Some tribal peoples
have raised concerns that this material represents their ancestral intellectual knowledge and that it was taken
from them without full disclosure of how it would be used. Many anthropologists published books and/or made
tenure at their universities based on such research. Meanwhile, little was done with the information to help the
tribal peoples it described, who were struggling under political and legal pressures to assimilate. In some
cases, tribal peoples have implemented research projects utilizing these manuscript collections that have the
explicit goal of helping their people with cultural recovery efforts.

One example of Indigenous peoples utilizing archive materials to their advantage is offered by Oregon’s
Coquille Indian Tribe, which made use of archival documents to successfully restore their tribe to federal
recognition in 1989 after the tribe was declared “terminated” by the federal government in 1954. Their
restoration bid was made difficult by the fact that the records of their tribal culture were collected in faraway
archives. Essential to the tribe’s success was George Wasson Jr., son of the aforementioned George Wasson who
was aided by Leonard Frachtenberg. Wasson Jr. designed and implemented an effort to collect copies of
anthropological manuscripts pertinent to the Coquille tribe from the Smithsonian Institution.

In 1995, 1997, and 2006, the Southwest Oregon Research Project—a project initiated by the Coquille Indian
Tribe, University of Oregon anthropologists, and students from western Oregon tribes—collected 150,000
pages of documents about the tribes of western Oregon from the Smithsonian Institution and the National
Archives. These materials have since become a major collection at the University of Oregon’s Knight Library
Archives, special collections division, and additional copies have been given to 17 regional tribes.

These projects are examples of the repatriation of intellectual knowledge to the tribes that the information was
collected from. Many libraries now have policies that allow concerned tribes to repatriate their intellectual
knowledge in the form of copies of collection materials for little or no cost. Recordings of songs represent a
particularly sensitive and special type of cultural artifact to many tribal people. Archives have historically not
been very attentive to the concerns of tribes regarding their collections. For more information, consult the
Protocols for Native American Archival Materials.

ETHNOGRAPHIC SKETCHES

Summers Collection and the Grand Ronde Tribe


by author David Lewis

The Summers Collection is a collection of more than 600 Native objects from the West Coast of the United
States, collected by the Reverend Robert Summers, an Episcopalian minister. A large portion of the collection,
some 300 objects, was collected from the Grand Ronde Indian Reservation, which is close to where Summers
lived in McMinnville, Oregon. In the 1870s, Summers would regularly visit the people of Grand Ronde and
purchase objects they had in their homes or were using. Most of these objects are woven baskets and trays
made in a traditional manner, many predating the formation of the reservation in 1856. Sometime in the

Access for free at openstax.org


2.6 • Collections 61

1890s, Summers passed his collection on to his associate Reverend Freer, who donated the collection to the
British Museum in 1900.

The collection has remained part of the British Museum collections since then. The value of this collection lies
not only in the objects and their unusually good preservation but also in the care Summers took to document
the people he purchased them from, their use, and their cultural background. It was unusual in early
anthropology for a collector to be so comprehensive in documenting material collections. Summers was likely
aided by his wife, who was a professional botanist and would have been meticulous in her work documenting
botanical collections.

In the 1990s, the Grand Ronde tribe became aware of the Summers Collection at the British Museum. In 1999,
representatives of the tribe visited the museum, viewed the collection, took photos of all objects related to the
tribes, and copied all the notes they could. Since then, the tribe has worked through a series of museum
curators to see if it would be possible to repatriate the collection to the Grand Ronde. The British Museum is
one of the largest repositories in the world, holding sacred and cultural objects from numerous nations, many
once part of Britain’s extensive colonial empire. The British Museum rarely allows repatriations, fearful that
allowing one to occur would set a precedent resulting in multiple other cultures submitting claims. Still,
curators of the North American collections have suggested that something could be worked out if there were a
book deal to help publicize their collections and significant enough publicity. In 2018, the Grand Ronde tribe
was able to negotiate the loan of some 16 objects from the collection. The objects were placed on display in the
new Chachalu Museum and Cultural Center in Grand Ronde. While there, the objects were studied by cultural
experts who focused on understanding how they were made and how they might be able to replicate the
techniques.

There are no protocols for international repatriation. The Grand Ronde tribe had to work diplomatically to
form negotiated agreements and establish a beneficial relationship with the British Museum. After more than
100 years of assimilation, many traditional skills had been lost to the Grand Ronde people. The opportunity to
regain some of this lost ancestral knowledge by studying these cultural goods is a rare gift.

MINI-FIELDWORK ACTIVITY

Participant Observation

When practicing participant observation, researchers immerse themselves in a cultural context and make
observations and notes about what occurs. This activity is structured to take place in a few hours and can be
accomplished in your community.

• Spend about an hour in a public place, such as a mall or store, coffeeshop, park, bus, train, or library, and
observe what people around you are doing. Take notes about their actions, interactions, clothing, foods,
mannerisms, and anything else that might seem interesting. Note characteristics and mannerisms
pertaining to culture, language, ethnicity, masculine and feminine roles, and age-related roles.
• Try not to be conspicuous, and do not record conversations unless they are spoken quite loudly so as not
to be intrusive. If anyone asks what you are doing, just explain that you have an assignment in a college
course to make an anonymous report on local culture.
• Return home and write a two-page reflective report on your research. In the report, give specifics of what
you witnessed, and analyze how you personally responded to different cultures or mannerisms. About
two-thirds of the report should be ethnographic reporting, and one-third should be analysis.
• Try to eliminate your personal bias or admit when you have one, and identify when you are basing your
analysis on personal opinions.
• Pay attention to the need to maintain the anonymity of your subjects as if this were an actual anthropology
fieldwork assignment. Do not identify people by name; instead, use pseudonyms.

As a final step, give a five-minute presentation about your experience that summarizes the high points of your
62 2 • Methods: Cultural and Archaeological

participant observation.

Suggested Readings
Boas, Franz. (1974) 1982. A Franz Boas Reader: The Shaping of American Anthropology, 1883–1911. Edited by
George W. Stocking Jr. Chicago: University of Chicago Press.

Boyd, Robert T., Kenneth M. Ames, and Tony A. Johnson, eds. 2013. Chinookan Peoples of the Lower Columbia.
Seattle: University of Washington Press.

Gross, Joan, ed. 2007. Teaching Oregon Native Languages. Corvallis: Oregon State University Press.

Kenoyer, Louis. 2017. My Life, by Louis Kenoyer: Reminiscences of a Grand Ronde Reservation Childhood.
Translated by Jedd Schrock and Henry Zenk. Corvallis: Oregon State University Press.

Konopinski, Natalie, ed. 2014. Doing Anthropological Research: A Practical Guide. New York: Routledge.

Lewis, David G. 2009. “Termination of the Confederated Tribes of the Grand Ronde Community of Oregon:
Politics, Community, Identity.” PhD diss., University of Oregon. http://hdl.handle.net/1794/10067.

Lewis, David G. 2015. “Natives in the Nation’s Archives: The Southwest Oregon Research Project.” Journal of
Western Archives 6 (1). https://doi.org/10.26077/e5e5-e0b1.

Sapir, Edward. (1949) 2021. Selected Writings of Edward Sapir in Language, Culture, and Personality. Edited by
David G. Mandelbaum. Berkeley: University of California Press.

Spradley, James P. (1980) 2016. Participant Observation. Long Grove, IL: Waveland Press.

Thwaites, Reuben Gold, ed. (1905) 2003. Original Journals of the Lewis and Clark Expedition, 1804–1806. Vol.
7. Madison: Wisconsin Historical Society. https://content.wisconsinhistory.org/digital/collection/aj/id/16212/
rec/7.

Access for free at openstax.org


2 • Key Terms 63

Key Terms
absolute dating methods (see also chronometric features cultural structures found at an
dating methods) dating methods that use archaeological site that are not movable or
physical and chemical properties of artifacts and portable, such as parts of a temple, altars, tombs,
structures modified by humans to establish their etc.
age without reference to other artifacts. For feminist anthropology an approach to
example, radiocarbon dating is used to date anthropology that seeks to transform research
organic materials generally up to 50,000 years methods and findings by engaging with more
old. diverse perspectives and using insights from
archaeobotanist a specialist who studies plants feminist theory.
and seeds appearing in an archaeology site. hypothesis a supposition that is subjected to
archaeological context the place where an object research in order to be proven or disproven
was originally found, along with other through data collection.
associations, such as the stratum it was found in, Indigenous anthropology the study of one’s own
specific features, and other objects associated culture or society using anthropological methods.
with it. The term has come to mean any application of
archaeological excavation the scientific process of Indigenous knowledge, perspectives, and
uncovering artifacts and other biological and scholarship in anthropology.
cultural remains in the historic and prehistoric institutional review board a university research
past of human-inhabited sites. committee that reviews biomedical or social
armchair anthropology a method of conducting science research proposals to determine if they
anthropological research without doing appropriately protect human participants,
fieldwork, relying instead on materials and informants, and subjects.
documents previously collected by others. interpretation the act of explaining the meaning of
artifacts objects that are portable and show something.
evidence of human cultural activity; for example, interview a method of research in which the
bones that show evidence of drawings sketched researcher asks questions of an informant to gain
on them, stone tools, pottery, etc. information about a person, society, or culture.
chronometric dating methods dating methods law of superposition the geological principle of
used to analyze various physical or chemical stratigraphy that assumes that materials,
characteristics of an artifact in order to assign a normally rock layers, found beneath other
date or range of dates for its production. materials are older that the materials on top.
cross-cutting relationship a principle in geology NAGPRA the Native American Graves Protection
and archaeology that suggests that a geologic or and Repatriation Act (1990), a US law that
cultural feature that cuts across another feature is protects human remains and cultural and
the more recently deposited of the two. ceremonial objects and artifacts from collection
dendrochronology an absolute dating technique and requires the return of such items already
that uses patterns of growth of tree rings and collected to the originating tribes. NAGPRA also
cross-dating to determine the approximate age of allows for the repatriation of the same materials
wood. from museums and other repositories.
ecofacts natural objects found at an archaeological naturalism an approach that seeks to understand
site, such as seeds, bone, shells, etc., that show no the world and the laws that govern it by direct
sign of human craftsmanship. observation of nature.
emic perspective viewing and attempting to open-ended in the context of anthropological
evaluate other peoples and cultures according to research, describes a research method whereby
the standards of those cultures; an “insider’s” the researcher allows informants to answer
point of view. questions without a limit in time or subject.
ethnology the study of differences and oral histories histories of previous events, moral
relationships between various peoples, societies, or ethical lessons, or stories of creation that are
and cultures. passed down by memorization. Many oral
etic (or ethnocentric) perspective viewing a histories are also called mythologies, legends,
culture from the perspective of an outsider texts, or folklore.
looking in. participant observation an anthropological
64 2 • Summary

research method in which the researcher enters a would go extinct and there would be nothing
cultural community and collects information further to study.
through observation of and participation in the scientific method a method of expanding
culture. knowledge by asking questions, creating a
primary context the context of an artifact, feature, hypothesis, collecting data, and presenting well-
or site that has not been disturbed since its reasoned findings based on evidence.
original deposition. secondary context the context of a cultural or
provenance the location of an artifact when it is natural objects that has been moved or disturbed
first found. The provenance is normally recorded from its original location and is thus no longer
when the artifact is in situ, or before it has been associated with its place of origin; for example, a
removed. burial that has been moved from its original
qualitative data nonnumerical data, such as location due to geological shifts or natural
language, feelings, or impressions, that is disaster.
normally collected when the researcher is at the seriation a relative dating method that places
research site. similar artifacts from the same area in a
radiocarbon dating a dating technique for organic chronological sequence.
substances that measures the decay of statistics the science of collecting and analyzing
radioactive carbon in the sample; also called numerical data in large quantities and inferring
carbon-14 (14C or C14) dating. This is the most proportions in a whole from those in a
widely used technique for dating organic artifacts representative sample, or the numerical data
between 50 and 60,000 years old. collected and analyzed in this manner.
relative dating describes methods of determining strata plural of stratum; in geology and
the relative order of past events through archaeology, distinct layers of deposited natural
comparisons of two or more artifacts without or archaeological material.
determining their absolute age; e.g., sample 1 is stratigraphic superposition a relative dating
older than sample 2 because sample 1 was found method that assumes that any cultural or natural
beneath sample 2. artifact that is found within a stratum, or that cuts
repatriation the process of returning human across two or more strata in a cross-cutting
remains, associated funerary objects, and relationship, is younger than the stratum itself.
ceremonial items to the originating culture. stratigraphy the process of identifying the order
research question a question that can be proved and relative positions of strata.
or disproved through research and observation. stratum singular of strata; one specific layer of
retesting the scientific practice of conducting deposited natural or archaeological material.
experiments or research more than once in order theory a supposition or a system of ideas intended
to determine if the findings are accurate. to explain something.
Retesting helps eliminate human and other errors three-dimensional collection a collection of
in testing and create a range of accuracy. objects or artifacts.
salvage anthropology a particular period in early typological sequence a set or group of objects
anthropological practices (1870s–1930s) during ordered according to their types.
which tribal cultures were subject to extreme zooarchaeologist an archaeologist who specializes
collecting from researchers. The practice in the identification of animal remains at an
occurred because of fears that Native cultures archaeological site.

Summary
Chapter 2 discusses how anthropologists gather include diverse settings such as one’s hometown (as
information. All of the subfields of anthropology in urban anthropology), the Internet (visual or
conduct fieldwork in some form to gather virtual anthropology), or collections in university
information, each subfield may use different archives and museums (ethnohistory or museum
methods of conducting research. The concept of anthropology). Research methods for cultural
working in “the field” was traditionally based on the anthropology and archaeology are covered in detail.
practice of traveling to distant regions to study other the chapter explores the issues that need to be
cultures within their native environmental contexts. considered when analyzing information gathered
In recent decades, “the field” has broadened to during research. This includes the biases of the

Access for free at openstax.org


2 • Critical Thinking Questions 65

anthropological researcher. Also covered is some of anthropological study and how fieldwork and
the history of the research methods used in methods have changed over time.

Critical Thinking Questions


1. Explain how conservation, as practiced in the anthropological inquiry? What does this
19th and early 20th centuries, attempted to approach offer to the field?
preserve animals, plants, and human cultures. 6. What rights do tribal peoples have when research
2. Describe salvage anthropology. Why was it is conducted on them?
practiced? What are some criticisms of this 7. What are the ethical responsibilities of
approach? anthropologists when conducting research?
3. What is different about anthropology as practiced What practices should anthropologists follow to
in the 19th century compared to the way it is be sure their research proceeds in a moral and
practiced today? ethical manner?
4. Why is it important to have multiple perspectives 8. Name and describe the most important pieces of
when describing human culture? How do government legislation in the United States
anthropologists gather these multiple pertaining to the rights of Indigenous peoples to
perspectives? cultural artifacts and knowledge.
5. What is a feminist anthropological approach to

Bibliography
Adams, William Mark. 2004. Against Extinction: The Story of Conservation. London: Earthscan.

Boas, Franz. (1974) 1982. A Franz Boas Reader: The Shaping of American Anthropology, 1883–1911. Edited by
George W. Stocking Jr. Chicago: University of Chicago Press.

Cole, Douglas. (1985) 1995. Captured Heritage: The Scramble for Northwest Coast Artifacts. Norman:
University of Oklahoma Press.

Hale, Horatio. 1846. Ethnography and Philology. Vol. 6, United States Exploring Expedition during the Years
1838, 1839, 1840, 1841, 1842, under the Command of Charles Wilkes, USN. Philadelphia: C. Sherman,
1844–1874.

Hodgen, Margaret T. (1964) 1971. Early Anthropology in the Sixteenth and Seventeenth Centuries.
Philadelphia: University of Pennsylvania Press.

Hymes, Dell. 1980. “What Is Ethnography?” In Language in Education: Ethnolinguistic Essays, 88–103.
Washington, DC: Center for Applied Linguistics.

Morgan, Lewis Henry. 1851. League of the Ho-dé-no-sau-nee, or Iroquois. Rochester, NY: Sage & Brother.
https://hdl.handle.net/2027/nyp.33433081750949.

Stocking, George W., Jr. 1966. “Franz Boas and the Culture Concept in Historical Perspective.” American
Anthropologist 68 (4): 867–882.
66 2 • Bibliography

Access for free at openstax.org


CHAPTER 3
Culture Concept Theory: Theories of
Cultural Change

Figure 3.1 Now and in the past, people have called many different types of dwelling home. Top left, tent on K
Street in Washington DC; top right, Puye cliff dwellings near Espanola, New Mexico; bottom left, a water village,
Brunei, Indonesia; bottom right the International Space Station (credit: top left “K Street” by Daniel Lobo/flickr,
Public Domain; top right “Puye Cliff Dwellings” by BFS Man/flickr, CC BY 2.0; bottom left “The water village. Burnei”
by Bernard Spragg. NZ/flickr, Public Domain; bottom right “The International Space Station after arrival of ISS Roll
Out Solar Arrays” by NASA/NASA.gov, Public Domain)

CHAPTER OUTLINE
3.1 The Homeyness of Culture
3.2 The Winkiness of Culture
3.3 The Elements of Culture
3.4 The Aggregates of Culture
3.5 Modes of Cultural Analysis
3.6 The Paradoxes of Culture

INTRODUCTION Though all humans have a set of basic needs, we meet those needs in very different ways in
response to environmental conditions and social circumstances. For example, consider the basic human need
for shelter. In places prone to flooding, people often build their houses on stilts, constructing patios and
walkways to connect their houses together. In mountainous areas, people sometimes carve their houses into
cliffsides. In societies with extreme inequality, some people live in luxury highrise apartments side-by-side
68 3 • Culture Concept Theory: Theories of Cultural Change

with people who pitch their tents on the sidewalk. Humans have even constructed a complex dwelling adapted
to the conditions of space, the International Space Center.

Similarly, humans have a wide range of solutions to human needs for clothing, food, family life, health, and
social order. In each society, the various solutions combine in a complex totality called culture. In this chapter,
we explore the concept of culture, what it is and how to study it. Taking the need for shelter as a central
example, we will see how culture is created and how it changes. We will learn about how different elements of
culture interact with one another. As culture is a central concept in anthropology, our understanding of culture
will guide our exploration of human lifeways throughout this textbook.

3.1 The Homeyness of Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Explain the importance of culture to the concept of home.
• Identify the centrality of culture in the discipline of anthropology.
• Describe how each of the four fields deploys the concept of culture.
• Explain why culture feels familiar and “homey.”

FIGURE 3.2 The floor plan of a typical 21st-century middle-class American house consists of many individual
rooms, including three to four bedrooms, a large kitchen, a family room, and an attached garage. This floor plan
depicts such a home. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

What place do you call home? For some people, home is a large, angular structure made of wood or brick, fixed
on a permanent foundation of concrete, and rigged with systems to provide running water, electricity, and
temperature control. Such houses have separate rooms for distinct activities, such as sleeping, bathing, eating,
and socializing. Often, one bedroom is larger than the others and connected to its own bathroom. This is the
“primary bedroom,” designed to accommodate a married couple while their children sleep in smaller
bedrooms. The room for cooking (the kitchen) used to be separated from the room where people socialized
(the living room or great room), as it was assumed that one person (the wife) would cook in the kitchen while
another person (the husband) relaxed alone or with company in the living room. More recently, open-concept
architecture has eliminated the wall separating the kitchen from the living room, as adults often cook together
or socialize as one cooks and the other relaxes.

Access for free at openstax.org


3.1 • The Homeyness of Culture 69

In the 1960s, French scholar Pierre Bourdieu (1970) analyzed a typical house of a Kabyle family in northern
Algeria. Traditional Kabyle houses were rectangular buildings made of stone and clay with tiled roofs. Inside, a
waist-high dividing wall marked off one-third of the house. This marked-off section, set lower than the rest of
the house and paved with flagstones, was the stable, where animals were kept at night. A farming people, the
Kabyle kept oxen, cows, donkeys, and mules. Above the stable was a loft where women and children often slept,
though arrangements for sleeping and marital sex tended to vary.

FIGURE 3.3 These houses in Norther Algeria, built by the Kabyle people, are constructed of stone, and include
open space for both animals and human inhabitants under a shared roof. (credit: PhR610/flickr, CC BY 2.0)

The floor of the larger section of the house was higher and paved with a layer of black clay and cow dung that
women polished with a stone. This part was reserved for human use. In this larger, elevated section, a large
weaving loom sat against the wall opposite the door. Facing east, this wall with the loom received the most light
in the house. Guests and brides were seated here, as it was considered the nicest part of the house. Opposite
the dividing wall in the larger section was the hearth, surrounded by cooking tools, lamps, and jars of edible
grain. With the loom and the hearth, the main area of human activity in the house was associated with the
work of women. Bourdieu explained that men were expected to remain outside the house from dawn to
evening, working in the fields and associating with other men in public spaces. Women were supposed to
remain in the home.

In Bourdieu’s analysis, the Kabyle house was divided into two realms: a dark, low realm associated with
animals and natural activities (sleeping, sex, childbirth, and death) and a lighter, higher realm associated with
humans and cultural activities (weaving, cooking, brides, and guests).

Humans all over the world require a place to gather, work, socialize, and sleep. Some have Western-style
houses, while others have compounds. Some live in tents made of wooden beams and covered with animal
skins or cloth, in caves hollowed out of sandstone or volcanic rock, or in wooden structures built on stilts or in
trees to avoid floods and predators. While these different forms of home are all designed to perform a common
function as human living spaces, they are distinctively shaped by local environments and lifeways. Houses are
most commonly built with locally available materials and designed to protect against local climatic conditions
and predators. Over generations, people develop distinctive technologies to transform available materials into
durable and functional homes. Different forms of family, different gender roles and relations, and different
70 3 • Culture Concept Theory: Theories of Cultural Change

everyday activities determine the organization of space in these different homes. Dominant ideas about work,
gender, marriage, parenting, hospitality, and status all shape the places we call home.

Home, then, involves a combination of materials, technologies, social relationships, everyday practices, deeply
held values, and shared ideas. In every culture, these features are uniquely combined to produce distinctive
versions of home. Other combinations of features produce distinctive versions of clothing, food, work, and
health. Growing up in a particular social group, a person learns these ways of living, eating, working, and so on
and comes to consider them normal and natural. Anthropologists have a word for such integrated
combinations of social and environmental features, and that word is culture. The ways of your culture are
familiar to you, often so deeply ingrained that they come naturally. Culture itself feels like home.

All four fields of anthropology are devoted to understanding human culture. Biological anthropologists are
often interested in the emergence of culture in the course of human biological evolution. Archaeologists use
material artifacts as keys to understanding the technologies, social practices, and ideas of ancient peoples.
Cultural anthropologists often use participant observation to understand how the various features of culture fit
together in contemporary societies. Linguistic anthropologists are interested in how language shapes and is
shaped by other features in the constellation of culture.

This chapter explores culture as a central concept in anthropology. We examine what distinguishes culture
from other aspects of human experience and activity. In an effort to organize the vast array of things included
in culture, we divide culture into three levels and consider how those levels fit together holistically—and what
happens when they don’t. Finally, we identify a set of contradictions built into the concept of culture and see
how those contradictions illuminate the nature of human social life.

3.2 The Winkiness of Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Provide E. B. Tylor’s definition of culture.
• Distinguish natural behavior from cultural behavior.
• Describe deliberate and nondeliberate ways that people acquire culture.
• Explain how biological processes can be shaped by culture.

In the last section, we referred to culture as a combination of materials, technologies, social relationships,
everyday practices, deeply held values, and shared ideas. Nineteenth-century British anthropologist Edward
Burnett Tylor defined culture as “that complex whole which includes knowledge, belief, art, morals, law,
custom, and any other capabilities and habits acquired by man as a member of society” (1873, 1:1). That’s a lot
to include in one concept! If all of that is culture, then what about human experience and activity is not
culture?

Consider this scenario. A student comes to class one day, and the instructor says, “I’ve decided that you’re all a
bunch of failures and I’m flunking the entire class.” Imagine then that the instructor simply stands there after
that announcement, blinking calmly as the class erupts in protest.

Now imagine that same scenario with one very slight difference. The instructor announces, “I’ve decided that
you’re all a bunch of failures and I’m flunking the entire class.” Then, as the class erupts in protest, the
instructor calmly blinks one eye, leaving the other eye open.

Access for free at openstax.org


3.2 • The Winkiness of Culture 71

FIGURE 3.4 Would you take this woman seriously? In American culture, winking, related to the normal biological
function of blinking, takes on special meaning in social interactions. (credit: Motion Picture News/Wikimedia
Commons, Public Domain)

What just happened there? Blinking is a biological compulsion common to humans everywhere. Humans blink
to keep eyes hydrated and clear of debris. Humans are born knowing how to blink; nobody has to teach us. On
average, humans blink 15 to 20 times every minute. Without realizing it, people are necessarily blinking
throughout every conversation, every social interaction, every activity during the day. The people we talk to
and interact with are also blinking constantly, so often that everyone is accustomed to ignoring it. Blinking
does not affect the perceived meaning of speech or actions.

But if someone deliberately blinks one eye, leaving the other one open, that’s a completely different matter. In
fact, leaving one eye open makes a blink a wink. Winking is not a biological necessity. Humans are not born
knowing to how to wink, and it takes some practice to learn how to do it. Because it requires deliberate effort
and people are not constantly doing it, winking can acquire special meaning in social interactions. In
American culture (and many others), a wink often indicates that someone is joking around and that whatever
they’ve just said or done should not be taken seriously. Of course, a wink can mean different things in different
societies. Moreover, a wink can mean different things in the same society. If someone on a date takes their
companion’s hand and gives a cute little wink, the person may have reason to hope the winker is not just joking
around.

American cultural anthropologist Clifford Geertz (1973) used the example of winking to illustrate two
important aspects of culture. First, culture is learned. Innate human behaviors—that is, behaviors that people
are born with—are biological, not cultural. Blinking is biological. Acquired human behaviors—that is, behaviors
that people are taught—are cultural. Winking is cultural. This means that cultural behaviors are not genetically
inherited from generation to generation but must be passed down from older members of a society to younger
members. This process, as you’ll recall from What is Anthropology? is called enculturation.

Some aspects of enculturation are deliberate and systematic, such as learning the rules of written punctuation
in a language. At some point in an English speaker’s childhood, someone explicitly told them the difference
between a question mark and an exclamation point. Most likely, they learned this distinction in school, a
fundamental institution of enculturation in many societies. Religious institutions are another common force of
72 3 • Culture Concept Theory: Theories of Cultural Change

enculturation, providing explicit instruction in cultural rules of morality and social interaction.
Extracurricular activities such as sports, dance, and music lessons also teach children cultural rules and
norms.

While a great deal of very important cultural content is deliberately conveyed in these systematic contexts, the
greater part of culture is acquired unconsciously by happenstance—that is, nobody planned to teach it, and no
one made an effort to consciously try to learn it. By virtue of growing up in a culture, children learn what
certain actions and objects mean, how their society operates, and what the rules are for appropriate behavior.

Going back to the cultural notion of home, did anyone ever explain to you why your childhood home was
structured in a certain way? Did anyone ever point out the cultural assumptions about gender and family built
into your house? Probably not. Now, imagine that you were taken away from your parents as a baby and
adopted by a family far away, with a very different way of life situated in a very different environment. With
your adoptive family, you might have been raised in a very different kind of home. Growing up, your everyday
habits, activities, and expectations would have been shaped by the setup of that home. Living in that house, you
would have wordlessly absorbed a set of assumptions about family, gender, work, leisure, hospitality, and
property. And all of it would seem quite natural to you.

Many forms of culture are passed down through a combination of deliberate and unconscious processes.
Perhaps when you were a child, someone told you what a wink was and showed you how to accomplish one; or
perhaps you just witnessed a few winks, figured out what they meant from their contexts, and then learned
how to accomplish one through trial and error. Geertz pointed out that there are two important aspects to
winking: the meaning and the action. As both are learned, both are cultural. But perhaps more importantly,
both the standardized action of winking and the assumed meaning of this action are commonly known among
members of a group. That is, culture is shared.

Consider another aspect of human biology: dreaming. People in all societies dream, and no one has to teach
them how to do it. Dreaming is biologically innate and spontaneously performed. Biological researchers
hypothesize that dreaming helps the human brain process daily stimuli and convert recent experiences into
long-term memories. As a biological necessity for brain health, dreaming is natural, not cultural.

But why do people dream in stories? And why are those stories so often confusing, even troubling? In many
cultures, people are perplexed by their dreams, never really knowing what the objects and situations they
dream about are meant to indicate—or if they have any meaning at all. In other cultures, however, dreams are
recognized as arenas of spiritual communication with supernatural beings. In Ojibwa culture, young people
are encouraged to fast for up to a week in order to bring on special visionary dreams (Hallowell 1992; Peters-
Golden 2002, 188–189). In such dreams, a young person may be approached by a guardian spirit who imparts
knowledge for successful hunting, warfare, or medicine. People are discouraged from discussing the meaning
of these dreams, but young people are taught to expect and anticipate this kind of dream, and they know how
to interpret the content of such dreams without discussion. The widely shared ability to dream such dreams
and the shared knowledge to understand their content makes dreaming profoundly cultural among the
Ojibwa.

Summing up, when an element of human experience or behavior is learned and shared, we know it is an
aspect of culture. That delineates the concept of culture to some degree. However, the variety of things that are
learned and shared by humans in groups is still quite enormous, as indicated by Tylor’s rambling list
(knowledge, belief, art, morals, law, custom, etc.). Instead of thinking of culture as one vast hodgepodge of
things, it’s helpful to break that hodgepodge into three basic elements. These basic elements of culture are
understood to come together in larger combinations, or aggregates.

Access for free at openstax.org


3.3 • The Elements of Culture 73

3.3 The Elements of Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Define the concept of material culture and provide examples of material culture.
• Provide a detailed example of cultural practices.
• Explain how cultural frames orient our experiences and actions.
• Describe how norms and values are threaded through culture.
• Explain how ideologies and worldviews shape our perception of the world around us.

The complex whole of culture can be broken down into three categories: what we make, what we do, and what
we think. The boundaries separating these categories are somewhat artificial because so much of cultural life
involves all of these things at once. However, it’s useful to start with the basic building blocks of culture, then
see how those blocks can be put together to produce more complex structures.

Culture Is What We Make


Museums are buildings where objects of historical, artistic, scientific, or cultural interest are displayed. The
Smithsonian’s National Museum of the American Indian has one of the world’s largest collections of Native
artifacts, including many two- and three-dimensional objects such as baskets, pottery, and preserved
specimens representative of the lives of Native populations from all areas of the country.

FIGURE 3.5 This basket, woven by Kucadikadi (Mono Lake Paiute) artist Lucy Telles, is an excellent example of the
art of basket weaving. Telles, whose work was done in the early part of the twentieth century, is widely admired for
her use of color and innovative designs. (credit: “Mono Lake Paiute Basket” by Ernest Amoroso, National Museum of
the American Indian/Wikimedia Commons, Public Domain)

People living in groups learn to craft the things they need in order to make a living in their environment. Early
human ancestors learned how to make sharp blades useful for processing meat. They shared their knowledge
of toolmaking in groups, passing those skills down to younger generations. Objects that are made and used by
humans in group contexts are called material culture. All of the tools developed by early hominins (blades,
arrows, axes, etc.) are examples of material culture. All of the artifacts discovered by archaeologists (buildings,
pottery, beads, etc.) are examples of material culture. The specialized knowledge and skills used for making
material culture are called technology. Today, the word technology is often used to refer to electronic devices
such as smartphones and computers. For anthropologists, both smartphones and obsidian blades are forms of
material culture produced through specialized technologies. That is, technology refers to the knowledge and
skills required to make blades, phones, and other objects of material culture.

Material culture is not just found in museums, of course. Material culture is all around. All of the furniture,
appliances, books, dishes, and pictures on the walls in a typical American home are elements of material
culture, and they reveal a great deal about the whole way of life of a society.

Consider the toothbrush. It would be possible for people to clean their teeth with a found object such as a twig
74 3 • Culture Concept Theory: Theories of Cultural Change

or leaf, or even with a finger. Ancient peoples often used a special chew stick, a twig with a frayed end. The
bristled toothbrush was invented in 15th-century China and spread across Europe and into the United States,
where it began to be mass produced in the late 19th century. Drugstores now feature many styles of toothbrush
with an array of special features. Specialized teams design, manufacture, and market this wide variety of
toothbrushes to consumers. Parents buy toothbrushes for their children and teach them the conventional
techniques for brushing their teeth (little circles, two minutes, etc.). As adults, people often isolate themselves
in a special room to brush their teeth in privacy. Even so, toothbrushing is a profoundly social act, relying on
shared knowledge and observance of social norms for hygiene and health.

Trees, rocks, microbes, and planets are all material objects, but they are not material culture unless they are
made and used by humans in group contexts. For instance, a tree growing in a natural forest is not an object of
material culture. However, an apple tree can be material culture if it is planted by a farmer in an orchard
designed to produce fruit for human consumption. A microbe can be material culture if it is manufactured to
improve human digestion or genetically engineered to fight cancer.

FIGURE 3.6 This rock is on display in the British Museum. While a rock is not in and of itself material culture, this
rock, which carries special meaning for those who view it, is. (credit: Archaeomoonwalker/Wikimedia Commons, CC
BY 3.0)

On display in the Cleveland Museum of Natural History is a gray rock. This rock was simply found by humans
and never shaped for any particular use. Sitting there in the museum, it has no specific purpose other than to
serve as an object of popular contemplation. Though it is a fairly unremarkable lump of basalt, thousands of
people stop to gaze at this rock, reading the sign that describes how it was obtained, marveling at its presence
there in the museum. Why? What’s so interesting about a rock? This particular rock was collected by
astronauts on the Apollo 12 mission to the moon. The rock serves as evidence of this magnificent feat of
scientific engineering and a source of great pride to the culture that accomplished such a mission. We go to
museums to view the items on display there, but clearly, the human activities surrounding those objects are
what make them interesting to us. That is, culture is not just material objects—it’s also what we do and what we
think.

Culture Is What We Do
Ahmed is a carpet seller in the Istanbul Grand Bazaar. Every day, people from all over the world come into his
stall to examine, and sometimes buy, the carpets in his inventory. Anthropologist Patricia Scalco (2019) met
Ahmed while she was conducting research on market exchange in Istanbul. She carefully observed the set of
sales strategies he had crafted to respond to customer desires and knowledge. When anyone pauses at the
entrance, Ahmed greets the potential customer and ushers the person into his stall. Bringing out a silver
platter, Ahmed offers the customer a cup of tea, a welcoming gesture. As the customer browses, Ahmed

Access for free at openstax.org


3.3 • The Elements of Culture 75

initiates a carefully constructed conversation designed to determine what sort of person this customer is, what
they are looking for, and what they really know (and do not know) about carpets. He pulls out various carpets
from the stacks, unfurling them as he describes their distinctive qualities. Ahmed identifies this interaction as
a sort of game he must play with his customers. European tourists in this Turkish marketplace are often
inspired by the desire for handmade traditional crafts made by local rural ethnic groups such as the Kurds.
These days, however, most carpets sold in the Istanbul market are industrially produced in Pakistan, India, and
China. However, in his many years of selling carpets, Ahmed has learned that he must play to Western
orientalist fantasies, weaving a distinctive story around the origins and manufacture of a carpet, in order to
win a sale. Like other merchants in this market, Ahmed has a family to support, and he cannot afford to openly
contradict the knowledge and desires of his customers.

Centered on the material culture of carpets, Ahmed’s work illustrates the importance of what people do and
what they think in the making of cultural life. What people do and what they think are nonmaterial elements of
culture. In his everyday interactions with customers, Ahmed has developed a set of habitual practices
involving gesture and speech. Anthropologists use the term cultural practices to refer to this form of culture.
Routine speech communicates meanings and values (such as the “authenticity” of a carpet), while routine
action organizes social events (such as, hopefully, a sale). People from all walks of life develop similar
combinations of habitual action and speech that constitute the everyday culture of people in those
circumstances.

What do you do in the morning to get ready for the day? That is cultural practice. What do you do when
someone comes over to your house? That is cultural practice. What do you do when you’re hungry? That is
cultural practice. Some cultural anthropologists focus on these everyday practices as keys to understanding
culture, while others are more interested in special events such as ceremonies and festivals.

For instance, Carnival in Brazil is an annual festival of music and dance held every year to mark the beginning
of the Catholic season of Lent. Parades of costumed dancers throng the streets of many cities, interacting with
the audience and attracting crowds of followers. Cultural anthropologist Kenneth Williamson (2012) studied
Carnival in Salvador, Bahia, in the north of Brazil. While Brazilian Carnival is framed as a national celebration,
Williamson found that Carnival in the poorer and largely Black city of Salvador is distinctively animated by the
politics of race. Local Carnival dance groups incorporate Black forms of movement such as capoeira, a
combination of dance and martial art techniques created by Brazilian enslaved peoples. Forms of music and
religion originating in Africa also contribute to the distinctiveness of Salvadoran Carnival. Carnival has
become increasingly commercialized as a tourist attraction in Salvador, bringing in Black and White tourists
alike. Black Brazilian activists complain that forms of Black culture are being appropriated and exploited as
forms of cultural leisure with little understanding of their deep cultural meanings as expressions of resistance
and survival. Meanwhile, most Black Salvadorans enjoy little benefit from the burgeoning tourist economy.

The practices of Turkish carpet merchants and Brazilian Carnival participants are both ways of doing culture,
every day and on special occasions. As we see in both examples, the materials and actions of culture are
infused with patterns of thought, some shared and some controversial. These ways of thinking constitute a
third element of culture.

Culture Is What We Think


Imagine that you are walking down the street and you see a building. You notice a mailbox next to a driveway.
You follow a little walkway lined with flowers to a front door. Below your feet, you find a mat that says,
“Welcome!” Peering through a window, you see a central room where two people are sitting on a couch, eating
chips, and watching television. Off to the side, there’s a hallway. You can barely see the stockinged feet of a
small person resting on a bed. A dog barks.

What kind of place is this? Are you sure? How do you know?

Now imagine you are walking down the street and see another building. There are neon lights in the front
window and a large paved area to the side. As you enter the front door, a little bell jingles and young woman in
a white blouse greets you from behind a long table. To one side of that table is a large black machine with
buttons and numbers on it. The young woman carries a small leather folder in her hand and gives you an
76 3 • Culture Concept Theory: Theories of Cultural Change

expectant smile. You look around to find a room full of people seated at tables of various sizes. Young people in
white tops and black pants are scurrying here and there, some carrying giant platters. You hear music in the
background. You smell something delicious.

What kind of place is this? How do you know?

In both scenarios, elements of material culture are combined with patterns of action and speech. In order to
make sense of these two scenarios, we must use shared ways of thinking about them. What we know about the
way of life in our society leads us to identify the first scenario as somebody’s home. What we know about the
circumstances of eating in public leads us to identify the second scenario as a restaurant.

These patterned, shared ways of making sense of situations are called cultural frames. Cultural frames tell
people where they are, what role they they play in that context, and what forms of behavior and speech are
expected and appropriate. There are cultural frames for places, times, events, and relationships. If a couple
have been dating for over a year, they probably use a cultural frame for romantic relationships to structure
their actions and expectations in that relationship. And if one of the romantic partners invites the other to
spend a holiday with their family, the invited person will probably summon a cultural frame for that holiday to
tell them what to expect and how to behave.

Cultural frames are complex cognitive models that incorporate various roles and actions patterned in space
and time. A cultural role is a conventionalized position held by a person or persons in a particular context or
situation. Sociocultural roles are associated with certain behaviors and actions. For example, “mother” is a
sociocultural role in the cultural frame of “family.” “Waiter” is a sociocultural role in the cultural frame of
“restaurant.” While these roles are found in many cultures, the actions and behaviors associated with them
vary significantly across cultural contexts.

In cultures that celebrate Mother’s Day, it is conventional to send one’s mother a card along with flowers and/or
a gift. Anyone who has ever been shopping for a Mother’s Day card has been bombarded with images and text
that convey the stereotypical behaviors and preferences associated with motherhood. Many Mother’s Day
cards feature pastel flower arrangements with birds, butterflies, and delicate calligraphy. The text lionizes the
emotional and material work of motherhood, praising the constant care and sacrifice of the good mother. In
return, the card promises eternal gratitude.

Access for free at openstax.org


3.3 • The Elements of Culture 77

FIGURE 3.7 This American Mother’s Day card from 1916 would still be considered appropriate today. The norm for
a Mother’s Day card in the United States has not changed much in over a century. (credit: Northern Pacific Railway/
Wikimedia Commons, Public Domain)

The behaviors and actions associated with a sociocultural role are collectively called a norm. Norms are not
necessarily “normal” in the sense that they represent the most common features and behaviors exhibited by
people in a certain role. Do all mothers prefer pastel flower motifs over, say, images of books or sports? Rather,
norms tend to be idealized, a fantasy about how people in a role behave—or how they should behave. Why do
we associate flowers, pastels, cursive, and self-sacrifice with motherhood?

The answer lies in another thinking element of culture: values. Cultural values are notions about what is good,
true, correct, appropriate, or beautiful. A certain mainstream way of thinking about motherhood indicates that
mothers should be delicate and feminine, concerned with beauty and decorum. Moreover, mothers should
nurture and sustain growth. What better way of conveying these notions than through the imagery of pastel
flower arrangements? Messages of gratitude describe the sort of behavior considered appropriate to mothers.
A “good” mother is a mother who puts her children at the center of her life at all times, neglecting her own
interests for the benefit of her family.

In any culture, norms indicate how people should behave, and values explain why they should behave that way.
For example, the norm for women in the 1950s was to get married and work in the home rather than have a job
in the public workforce. Not that all women did this, or even most. Many mothers, particularly women of color,
were obliged to work outside the home just to make a living for their families. Nonetheless, normative
depictions of women as housewives dominated media and public discourse in mid-20th-century America,
establishing this idealistic norm. Why were mothers supposed to stay at home? A set of “family values”
appointed fathers as the breadwinning heads of household, while mothers were relegated to serving men by
keeping house and caring for children. Thus, the values that came to be associated with motherhood were
subservience, self-sacrifice, gentleness, and nurture—the very values we see celebrated on Mother’s Day cards.

Norms and values can combine in larger models that depict how various social realms operate, such as the
family, the economy, the supernatural, and the political sphere. These models are known as ideologies. An
ideology identifies the entities, roles, behaviors, relationships, and processes in a particular realm as well as
the rationale behind the whole system. Take democracy, for instance. The political ideology of democracy
78 3 • Culture Concept Theory: Theories of Cultural Change

envisions a society of equal individual citizens who each cast a vote on proposals for government action. The
majority vote wins. The essential roles in this ideology are citizen voters and government. The essential
actions are voting and government action. The rationale is that government should obey the wishes of the
citizenry.

Is this how democracy really works, though? What about the influence of powerful organizations such as the
media and large corporations? Moreover, in most democracies, people do not vote directly on government
policies but rather elect representatives, who craft laws and then vote on those laws themselves. Those
representatives are accountable to citizens through the process of voting, but they are also strongly influenced
by lobbyists representing business interests and the campaign donations of wealthy individuals and groups.
Obviously, this ideology is a simplification of the way any democratic system really works. Ideologies are
always partial, foregrounding the perspectives of some people in society while obscuring the perspectives of
others.

A worldview is a very broad ideology that shapes how the members of a culture generally view the world and
their place in it. Worldviews tend to span several realms, including religion, economics, and politics. A
worldview provides an overarching model for the purpose and process of social life, depicting “how the world
works.” Many West African cultures, for instance, are shaped by a worldview that identifies the rationale of
society in the accumulation and distribution of material goods in extended families, communities, and the
nation as a whole. People rise to leadership through their ability to accumulate wealth, but they are strongly
obligated to distribute that wealth through their extended families and communities by funding the education
and business ventures of family members and helping those in need. Beyond the family, the actions of political
and business leaders are shaped by this worldview as well. A political leader is expected to support the
generation of wealth while also making sure that the benefits are spread through the community. Moreover,
leaders are expected to maintain relationships with departed ancestors who watch over their descendants.
Through periodic rituals and offerings, leaders petition ancestors to bless their families and communities with
prosperity and good fortune.

3.4 The Aggregates of Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Explain how elements of culture combine in aggregates.
• Give three detailed examples of cultural symbols.
• Explain how symbols are embedded in rituals.
• Describe how social structures organize important cultural processes.

An aggregate is a combination of elements. What we make, what we do, and what we think all combine in larger
aggregates of culture. For instance, it’s pretty clear that toothbrushes, moon rocks, restaurants, and Mother’s
Day cards must be understood as aggregates of material objects, practices, and ideas. In order to fully
understand the toothbrush as a cultural object, we must examine not only its design and production but also
how people use toothbrushes and why they use them. A set of routine practices surround our cultural objects
(brushing), and those practices are supported by cultural ideas (hygiene).

Symbols
A symbol is an object, image, gesture, vocalization, or event conventionally associated with a particular
meaning. anthropologist Jennifer Hasty was conducting fieldwork in Ghana during an election year, she
noticed that the posters and pamphlets of one politician featured a broom. Confused, she asked a friend why a
male politician would choose a humble domestic tool associated with women’s work as his political motif.
Making a sweeping motion with her hands, she explained, “Because he is promising to sweep away all the
corruption.” Turns out, he was the not the first to use this symbol. Over time, the broom has come to acquire
political meaning as a symbolic anti-corruption tool in Ghana.

Access for free at openstax.org


3.4 • The Aggregates of Culture 79

FIGURE 3.8 Featured in political posters and pamphlets, brooms like these have taken on special meaning in
Ghanian politics, where they are understood as a symbol of a politician’s intent to sweep away corruption. (credit:
“Handmade Brooms at Granville Island Broom Co.” by Ruth Hartnup/flickr, CC BY 2.0)

Colors, shapes, gestures, animals, plants—all of these commonly acquire specific cultural meaning. For a
Hindu wedding, a bride typically wears a bright red sari, as red is an auspicious color associated with change,
passion, and prosperity. White, on the other hand, is typically worn to Indian funerals.

Symbols are useful cultural aggregates because they provide a kind of shorthand for expressing complex ideas.
Consider the American bumper sticker shown in Figure 3.9.

FIGURE 3.9 This popular American bumper sticker incorporates a variety of religious and social symbols. (credit:
“Coexistence” by Rusty Clark/flickr, CC BY 2.0)

Combining symbols from Islam, Judaism, Taoism, Christianity, paganism, women, men, and the peace
movement, this sticker aims to promote multicultural diversity. Rather than listing the various religions,
identities, and ideologies and describing the conflicts among them, the message simply incorporates their
symbols into a word urging mutual tolerance.

Although symbols have conventional meanings, they can mean different things in different contexts or to
different people. Although the intended meaning of the above bumper sticker is diversity, some people
80 3 • Culture Concept Theory: Theories of Cultural Change

interpret it as an emblem of radical atheism. In the wake of the 2016 presidential election, some Americans
started wearing safety pins to show their solidarity with LGBTQIA+ people, people of color, and others who had
become targets of post-election harassment. For some, however, the safety pin symbolized pretentiousness
and hypocrisy.

Ritual
Combining objects, actions, and meanings, ritual is a special kind of repeated, patterned action conventionally
associated with a particular meaning. Rituals incorporate symbols and roles along with routinized activities
such as gestures, music, and movement. Many rituals are performed by specialists in group settings to
accomplish specific group or individual goals. Rituals bring together symbols, practices, and worldviews.

Consider this popular American ritual. On the first Sunday in February, many Americans gather in each other’s
homes to watch the annual championship game of the National Football League (NFL) on television. So
widespread is this practice that stores are nearly empty and many Christian churches cancel afternoon and
evening activities. As a whole, the ritual consists of many roles and relationships as well as patterned actions
and conventional meanings. At the heart of the action are the two teams competing against one another in a
chaotic game featuring an oddly shaped ball carried forward in campaigns of full-frontal assault across a
carefully marked field. The players are surrounded by referees, coaches, camera people, and cheerleaders,
each group having a strategic role in the action. Surrounding the field are commentators who interpret and
contextualize, giving meaning to the actions of the game. At home, some people watch the game closely,
exclaiming with joy or disappointment and commenting on the comments of the commentators. Other people
socialize with one another, watching the game intermittently. Vast amounts of food and drink are consumed by
Americans on Super Bowl Sunday. Typical foods include potato chips, dips, barbecued chicken wings, and
pizza. Beer is the beverage of choice for this occasion. An event celebrating competition, spectatorship, and
consumption, Super Bowl Sunday is an effective ritual for reinforcing dominant values in a society structured
by corporate capitalism. Notions of gender, race, and class are threaded through the various levels of play and
consumption as well.

In the Akan communities of central and southern Ghana, in West Africa, leaders perform a ritual called Adae
that uses important cultural symbols and reinforces cultural commitments to authority, ancestors, and shared
prosperity. In the Akan society, people are given special wooden stools to mark certain stages in life, such as
puberty and marriage. A person’s stool is said to contain the personal power of the owner, symbolizing the life
essence of that person.

FIGURE 3.10 This stool is more than just a place to sit down. In the Akan society which created it, it is understood
to represent the personal power and life essence of the person it was given to. (credit: “Stool (Dwa)” by Museum
Expedition 1922, Robert B. Woodward Memorial Fund/Wikimedia Commons, CC BY 3.0)

When an eminent person dies, that person’s stool is enshrined in a special shed called a stool house, or

Access for free at openstax.org


3.4 • The Aggregates of Culture 81

nkonuafieso. Twice every 42 days (once on a Wednesday and once on a Sunday), a community leader makes a
procession to the stool house of the ancestral leaders of the community. Entering the stool house, the leader
must remove their sandals and lower the cloth worn draped around their shoulders, symbolizing their
humility and respect for the ancestors. Then the leader greets the ancestral leaders one by one, making
offerings of drink and food and asking for blessings and prosperity for the community.

Special rituals called rites of passage are used to mark the movement of a person from one social status to
another. Naming ceremonies, puberty rites, weddings, and funerals are all common rites of passage.
Anthropologist Arnold van Gennep (1960) identified three stages in rites of passage: separation, transition,
and incorporation. In the first phase, separation, individuals, or groups are taken out of their everyday social
context, leaving their original social status. In the second phase, transition, people exist in an in-between state
outside of conventional norms of dress and action. In this phase, people are often dressed in special costume,
made to engage in unusual behaviors, and taught special forms of secret knowledge. In the third phase, people
are brought back into society in a formal ceremony and introduced as subjects in a new social category.

Initiation rituals are a common rite of passage in many societies. In many African societies that practice
initiation, young people are gathered together in a group and taken to a special camp outside the town or
village. This constitutes the separation phase. In the next phase, transition, members of the group are often
dressed alike and made to follow a common set of rules and schedule of activities. They may be required to
perform unusual feats, such as eating strange foods. Their bodies may be scarified or tattooed. Elders give
them special knowledge essential to performing their future roles as women or men. For instance, girls may
learn explicit lessons about conception and childbirth. Finally, when the transition is complete, initiates are
returned to the town or village and presented as women or men. Often, the completion of initiation marks a
young woman as formally eligible for courtship and marriage.

Social Structure
The way a society is formally organized is called social structure. Typically, a society organizes a set of routine
activities and objects in space and time to accomplish a particular function, such as community decision-
making, the production and circulation of goods, or religious observance. Social structure is the framework for
those realms, designating when, where, how, and by whom these functions are accomplished. Social structures
combine material culture (such as buildings) with practices (such as meetings) and ideas (such as the rules
and procedures of those meetings).

Consider the social structure of community decision-making, or the political realm. In some societies,
community decisions are routinely made under the authority of a person inhabiting an inherited political
office, called a chief or king (such as the Akans, discussed in the last section). Chiefs often have a council of
community elders, the heads of local extended families. A chief is expected to consult with this council in all
community matters. Other groups in society may represent the interests of youth, women, farmers, or traders.
Each group will have its own leader who communicates directly with the local chief. Regular procedures
govern how issues are raised and discussed and how decisions are taken. Together, the groups, roles,
relationships, and procedures all constitute the social structure of the political system.

Rather than seeing social structure as fixed and immobile, some anthropologists emphasize that people
continually make and alter their social structures through everyday forms of interpretation, participation, and
resistance. These processes mean that social structures are always subject to a variety of forces in a constant
state of change.
82 3 • Culture Concept Theory: Theories of Cultural Change

3.5 Modes of Cultural Analysis


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Explain how evolutionary theories have been applied to the study of human culture.
• Identify two critiques of evolutionary approaches.
• Describe how anthropologists have studied the functionality of culture.
• Distinguish Malinowski’s functionalism from Radcliffe-Brown’s structural functionalism.
• Explain how ontological anthropology defines the study of reality.

Anthropologists have a number of ways of studying the elements and aggregates of culture. Some approaches
emphasize the development of a particular aspect of culture over time, while other approaches examine how
the different parts of culture fit together.

Evolution, Adaptation, and Historical Particularism


Some anthropologists are interested in the origins of human cultural forms and how these forms have changed
over long periods of time. Just as Charles Darwin applied the notion of evolution to explain how biological
species change over time, many 19th-century anthropologists used evolution to explain how cultures changed
over time. This approach is called cultural evolutionism. Like Darwin, these anthropologists believed that
simple forms evolved into more complex forms. Comparing different cultures of the world, they assigned the
ones they considered more rudimentary to earlier evolutionary stages, while the ones they considered more
complex were assigned to the more advanced stages. For example, British anthropologist Edward Tylor argued
that human culture evolved from savagery through barbarism to civilization. He identified savagery with
people who used gathering and hunting to meet their basic needs. The domestication of animals and plants
was associated with barbarism. Civilization resulted from more advanced forms of farming, trade, and
manufacturing as well as the development of the alphabet. Not surprisingly, British scholars identified their
own culture as highly civilized.

Elaborating on Tylor’s scheme, American anthropologist Lewis Henry Morgan subdivided each of these three
stages into an even more elaborate model and proposed a mechanism for moving from stage to stage. Morgan
focused on technology as the primary driver of cultural evolution. New and better ways of making things,
according to Morgan, resulted in new patterns of social practice and thought. Advanced technology was
associated with advanced civilization.

But is technology the only measure of cultural accomplishment, or even the best one? Members of societies in
which people gather and hunt for a living have vast stores of knowledge about their environments. Typically,
they can name hundreds of plant species and tell when and where to find each of them. Many hunters can
examine animal tracks to discern the species, sex, age, and condition of the animal as well as how long ago the
tracks were laid. People in these societies also actively sustain and nurture diversity in their environments,
careful to avoid depleting important resources. Is it really accurate to think of such cultures as simple? All
cultures are complex, though in different ways. Technology is highly valued in American culture, while
environmental knowledge and sustainability have historically been less valued. Is it any wonder that early
American anthropologists ranked other cultures according to one of their own most cherished values? Perhaps
people in more environmentally sustainable cultures might consider the United States to be an example of
environmental savagery.

Both Tylor and Morgan, like most anthropologists of their day, thought that all cultures passed through this
single set of stages in the march toward civilization. This kind of theory is called unilineal evolution.
Disagreeing with this way of thinking, anthropologists such as Franz Boas argued that there is no single line of
cultural evolution but that each culture changes according to its own unique historical trajectory. Moreover,
cultures evolve not in isolation but in constant interaction with one another. Rather than focusing on
technological changes within a culture, Boas highlighted the diffusion of material objects, practices, and ideas
among cultures in complex relations of trade, migration, and conquest.

Though theories of unilineal cultural evolution have been largely abandoned, some anthropologists are still

Access for free at openstax.org


3.5 • Modes of Cultural Analysis 83

interested in discovering regular patterns that might govern how human cultures change over long periods of
time. In the 1950s, American anthropologist Julian Steward developed an approach called cultural ecology,
recognizing the importance of environmental factors by focusing on how humans adapt to various
environments. Steward’s approach showed how humans in each environmental zone develop a set of core
cultural features that enable them to make a living. Central to each cultural core are ways of getting or making
all the resources necessary for human survival—in particular, food, clothing, and shelter. Similarly,
anthropologist Marvin Harris developed a theory called cultural materialism, arguing that technology and
economic factors are fundamental to culture, molding other features such as family life, religion, and politics.

Though recognizing the importance of cultural change, many anthropologists reject the notion that all cultures
change according to a general universal model, such as cultural materialism. Drawing from the Boasian notion
that each culture follows its own historical path; many cultural anthropologists analyze change in terms of
historical particularism. In this approach, contemporary processes are understood as products of the unique
combination of internal and external forces unfolding over time in a particular culture.

Functionalism
Rejecting the comparative unilineal models that assigned each culture to an evolutionary stage, a number of
cultural anthropologists developed a radically different approach that attempts to understand each
contemporary culture in its own terms. Functionalism seeks to understand the purpose of the elements and
aggregates of culture in the here and now.

Bronislaw Malinowski, an early proponent of this approach, argued that the function of culture is to meet
human needs. All humans need to satisfy the need for food, clothing, and shelter. The fundamental purpose of
culture is to provide a means of satisfying those needs. In the course of meeting those basic needs, humans in
all cultures develop a set of derived needs—that is, needs derived from the basic ones. Derived needs include
the need to organize work and distribute resources. Family structures and gender roles are examples of
cultural elements addressing these derived needs. Finally, cultures also address a set of integrative needs,
providing people with guiding values and purpose in life. Religion, law, and ideologies fulfill these integrative
needs. Malinowski sought to understand both the biological and psychological functions of culture.

At first glance, this approach may not seem all that different from evolutionary approaches that identify the
core set of cultural features devoted to human survival. What was so different in Malinowski’s approach was
his attempt to show that even so-called primitive societies had functionally complex cultural systems for
meeting the full array of human needs. Malinowski’s three-volume ethnography of the economics, religion, and
kinship of the Trobriand people of Papua New Guinea demonstrated this fact in striking and elaborate detail.

A second version of functionalism, advocated by British anthropologist Alfred R. Radcliffe-Brown, identified


the functions of various elements of culture in a slightly different way. Rather than looking for the way culture
satisfies biological or psychological needs, structural functionalism focused more on how the various
structures in society reinforce one another. Culture is not a random assortment of structural features but a set
of structures that fit together into a coherent whole. Common norms and values are threaded through the
family structure, the economy, the political system, and the religion of a culture. Structural functionalists
conceptualized culture as a kind of machine with many small parts all working in tandem to keep the machine
operating properly. While recognizing the value of this approach, contemporary anthropologists have
complicated the mechanistic model of culture by pointing out that the various elements of culture come into
conflict just as often as they reinforce one another. Although few anthropologists would now identify
themselves as structural functionalists, the holistic approach to culture as an integrated system is derived
from this important theoretical foundation.

Structuralism
In the previous paragraph, you learned about structural functionalism, an approach that marries
functionalism with social structure. In a different sense, the term structure can refer to patterns of thought
embedded in the culture of a people—that is, conceptual structure. French anthropologist Claude Lévi-Strauss
pioneered this approach, sometimes called French structuralism. Lévi-Strauss considered culture to be a
system of symbols that could be analyzed in the various realms of culture, including myths, religion, and
84 3 • Culture Concept Theory: Theories of Cultural Change

kinship. In these realms of culture, objects and people are organized into symbolic systems of classification,
often structured around binary oppositions. Binary oppositions are pairs of terms that are opposite in
meaning, such as light/dark, female/male, and good/evil. For example, kinship systems are varied and
complex, but they are fundamentally structured by oppositions such as male versus female, older versus
younger, and relation by blood versus relation by marriage. Lévi-Strauss examined myths as well, showing how
the characters and plots emphasize binary oppositions. Consider the many European folktales featuring an
evil stepmother (Cinderella, Sleeping Beauty), a character that combines the opposition of good versus evil
with the opposition of blood relation versus relation by marriage. Lévi-Strauss argued that myths operate as
public arenas for conceptually pondering and processing the fundamental categories and relations of a
culture.

Ontology
In recent decades, some cultural anthropologists have come to focus on the nature of reality, including but not
limited to human perspectives and experiences. Ontology is the study of the true nature of existence. In some
cultures, for instance, the social world consists not only of embodied persons but also of spirit beings, such as
ancestors and witches, who interact with people in mysterious ways. And in some cultures, people are not just
bodies but assemblages that include souls, spirits, characters, or fates. Ontological anthropology explores
how culture constructs our social and natural realities, what we consider real, and how we act on those
assumptions. Reaching beyond human realities, ontological anthropology also attempts to include nonhuman
perspectives, relationships, and forms of communication.

For instance, in his provocative ethnography How Forests Think (2013), anthropologist Eduardo Kohn
describes how the web of life in the Amazon rainforest consists of continual communication among plants,
animals, and humans. He examines how Amazonian peoples engage with dogs, spirits, the dead, pumas, rivers,
and even sounds. Humans and these nonhuman beings are both antagonistic and interdependent in this
interactive web. Predators and prey read one another’s behavior, interpreting intentions and motivations.
Kohn’s effort is to get beyond conventional modes of human thought and language to understand how humans
are embedded in nonhuman ecological realities.

PROFILES IN ANTHROPOLOGY

Dame Mary Douglas


1921–2007

Dame Mary Douglas (https://www.openstax.org/r/wiki_Mary_Douglas).

Personal History: Mary Douglas was born in San Remo, Italy; her British parents had stopped off on their way
home from Burma, where her father had been working as a colonial civil servant. As children, Mary and her
younger sister lived with their mother’s parents in England until they were old enough to be sent to Catholic
boarding school—a fairly common practice for the children of colonial officers. After the death of her mother
and her dearly loved maternal grandfather, young Mary found security in the order and routine of the convent
school (Lyons 2011). This respect for rules and order combined with a reverence for the Catholic Church to
shape her lifelong commitment to studying the sacred aspects of the social order.

Area of Anthropology: At Oxford, Douglas studied with the prominent structural functionalist E. E. Evans-
Pritchard. From him, she learned that African belief systems such as witchcraft were structured by an
underlying logic. In this approach, the goal of fieldwork is to examine oral forms of culture as well as ritual and
social practice in order to discern the underlying logic that governs culture as a whole. Douglass went to the
Kasai region of what was then the Belgian Congo, where she studied how the Lele people used animals in
practical and symbolic ways. She was particularly interested in a strange animal called the pangolin. Though a
mammal, the pangolin has scales and no teeth.

Access for free at openstax.org


3.5 • Modes of Cultural Analysis 85

FIGURE 3.11 This pangolin is classified as a mammal but has scales like a reptile or fish. Pangolin were considered
sacred to the Lele people, who did not classify them as a food animal. (credit: Official photographer of the U.S.
Embassy in Ghana/Wikimedia Commons, Public Domain)

Douglas described how the Lele observed a fundamental distinction between edible and inedible animals.
Animals who lived among humans, such as rats and domesticated chickens, were considered part of society
and therefore inedible (most of the time). Only wild animals were considered food. Pangolins are wild animals,
but the Lele did not eat them (usually). Why? Douglas argued that the weirdness of the pangolin made people
single it out for special consideration. Pangolins have scales like fish, but they live on land and climb trees.
They look vaguely reptilian, but they do not lay eggs, instead giving birth to live young. Rather than teeth, they
have long snouts that they use to vacuum up small insects. Thus, the pangolin defied the conventional
categories the Lele used for dividing up the animal world. This breach of categories made the pangolin both
repellent and sacred to the Lele. Members of a special fertility cult engaged in rituals in which they ate
pangolins to ingest the power of this anomalous animal.

As this examination of cultural categories and anomalies suggests, Douglas was also influenced by Claude
Lévi-Strauss and the approach of French structuralism. Like Lévi-Strauss, Douglas viewed culture as a
coherent system of categories that were expressed in oral culture and social practice.

Accomplishments in the Field: Following her work on the Lele people, Douglas went on to conduct a broadly
comparative study of objects, practices, and people that were considered ritually dangerous, subject to rules of
prohibition called taboos. She showed how the subjects of taboos are often “matter out of place” (Douglas 1966,
44), things that defy conventional categories for dividing up the social and natural world. In her most famous
work, Purity and Danger (1966), Douglas examines a wide range of taboos, such as rules against eating certain
foods or engaging in sex at certain times or with certain persons. She examines the set of social and dietary
rules established by ancient Hebrews, detailed in the book of Leviticus in the Old Testament. According to
these rules, the Jewish people were forbidden from eating pigs, shellfish, and certain wild animals. They were
not allowed to wear garments made of cloth that combined different fibers—such as, for example, a linen-
cotton blend. Men were prohibited from having sex with menstruating women. In fact, women were
considered so unclean during menstruation that anyone or anything that touched a menstruating woman
became contaminated for the rest of that day.

What do all of these prohibitions have in common? Douglas shows how each forbidden object or condition
produced discomfort because it transgressed conventional categories. Shellfish, for instance, are sea animals,
but they don’t have fins or scales, and many of them do not swim. Menstruation is blood loss, but it does not
indicate injury. Moreover, menstruation is hidden and connected to the dangerous states of pregnancy and
childbirth. In Hebrew law, menstruation itself was considered a dangerous and contaminating exception to the
purity of persons and objects.
86 3 • Culture Concept Theory: Theories of Cultural Change

In her later work, Douglas applied this style of analysis to a variety of other social phenomena, including
humor and trickster figures. She argued that humor functions as a release for thoughts and actions that might
threaten the social order. Whereas taboos regulate and prohibit interaction with dangerous objects, animals,
and people, humor seeks to sap them of their dangerous power by making light of them.

Importance of Her Work: After more than 25 years of teaching at the University of London, Douglas moved to
the United States, where she held positions at the Russell Sage Foundation and Northwestern University. She
continued to publish widely on such topics as consumerism, environmental risk, and decision-making in
bureaucracies. When she retired, she moved back to England. In 2006, she was made Dame Commander of the
Order of the British Empire. She died in 2007 at the age of 86.

3.6 The Paradoxes of Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Identify four paradoxes in the concept of culture.
• Define four mechanisms of cultural change.
• Provide a detailed example of the mobility of culture.
• Describe culture as an arena of argument and contest.
• Explain how members of a culture can have different versions of their shared culture.

As European immigrants settled in the western frontier of the United States, they faced the challenge of
reinventing the elements of culture familiar to them in very different environmental and social conditions.
Used to living in houses made of wooden planks or logs, they found themselves on vast plains with very few
trees. A common adaptation to this environmental limitation was to dig into a slope of earth to create a dugout
home with turf walls and roof.

FIGURE 3.12 This Nebraska home, photographed with cow on its roof in 1870, was constructed in the side of the
hill directly behind it. While such dugout homes were practical and functional, those who lived in them typically
strove to replace them with wood-frame houses, as symbols of wealth and achievement. (credit: Solomon D.
Butcher/Library of Congress, Public Domain)

While these homes were perfectly functional, many Euro-American settlers considered them dirty and
backward. When their farming ventures became prosperous, they often undertook the great expense of
importing wood from forested areas to build the kind of house familiar to them from life back east, either on
the East Coast of the United States or in the European countries they originally came from.

While conducting fieldwork in Lesotho in the 1980s, cultural anthropologist Jim Ferguson observed that

Access for free at openstax.org


3.6 • The Paradoxes of Culture 87

people who became prosperous often replaced their round homes made of mud and stone and thatched roofs
with rectangular ones featuring cement floors and galvanized steel roofs. While the round buildings were
functionally adapted to local conditions, made of local materials, cool on hot days, and warm in cool nights, the
rectangular ones heated up like ovens under the hot sun and were noisy in the rain. The materials were
imported and expensive. Talking to one man who was planning to replace his round house with a rectangular
one made of cement and steel, Ferguson suggested that local building methods and materials might be
superior to foreign ones.

Looking me carefully in the eye, he asked, “What kind of house does your father have, there in
America? ... Is it round?” No, I confessed; it was rectangular. “Does it have a grass roof?” No, it did not.
“Does it have cattle dung for a floor?” No. And then: “How many rooms does your father’s house
have?” ... I mumbled, “About ten, I think.” After pausing to let this sink in, he said only: “That is the
direction we would like to move in.” (Ferguson 2006, 18)

In both cases, for Euro-American settlers and Lesotho villagers, the idea of home is not a settled matter but
subject to the forces of environmental adaptation, functionality, social status, and ideological debate. Both
examples illustrate a set of tensions at the heart of the concept of culture. Originally, anthropologists studied
culture as a fairly stable and consensual set of features commonly embraced by the people of a certain
geographical area. In the course of the 20th century, however, anthropologists began to realize that this notion
of culture was misleading and incomplete. In the early 20th century, American anthropologist Franz Boas
argued that the elements of culture are highly mobile, diffusing through the cultural contacts of trade and
migration. Since the 1960s, cultural anthropologists have come to emphasize the controversial aspects of
culture: how people disagree and argue over the dominant values and practices of their societies. Much of this
controversy stems from the unevenness of culture within a society—how people in different social categories
and subgroups participate differently in their common culture, with different versions or perspectives on the
same cultural norms and practices.

Despite these forces of change and controversy, there is something durable and shared about culture, some set
of common elements that distinguishes the whole way of life of each society. Even as cultures change through
innovation and contact, they often hold on to some of their distinctive features. In the 1980s, some scholars
thought that increases in global trade, migration, and technology were transforming all the diverse societies of
the world into one uniform global monoculture. In the 2020s, we see that the opposite has happened. In many
parts of the world, we have seen a resurgence of cultural identities and explicit efforts to maintain, rehabilitate,
and reinvent forms of cultural heritage.

So riddled with contradictions is the concept of culture that some anthropologists have suggested ditching the
whole notion altogether and finding some other concept to bind together the four fields in their pursuit of
knowledge about humanity. Perhaps such an integrated understanding of humanity isn’t even possible.

Or maybe the contradictions of culture are the most illuminating aspects of the culture concept. Maybe those
contradictions are anthropology’s most important contribution to our understanding of humanity. This
textbook takes the latter approach. Culture is the whole way of life of a people subject to a set of contradictory
forces. These forces constitute four central paradoxes of culture.

Paradox 1: Culture Is Continuous, but It Changes


Cultural materials, practices, and ideas are handed down from older to younger members of a culture, giving
some degree of continuity to culture over time. However, many factors can intervene in this process of cultural
reproduction to subtly alter or dramatically change the elements and aggregates of culture. In some contexts,
younger people either fail to precisely learn the culture of their elders or deliberately reject those cultural
lessons. Through travel and trade, people learn about other ways of doing things, and they take these ideas
back to their own cultures, trying them out to see how they might improve their own ways of life. Accidents and
deliberate experimentation introduce new possibilities. People may simply get tired of doing things one way
over and over and thrill at some refreshing style or craze.

We can identify four main mechanisms of cultural change. These four mechanisms overlap and interact as the
history of a culture unfolds over time. Diffusion is the movement of an element of culture from one society to
88 3 • Culture Concept Theory: Theories of Cultural Change

another, often through migration or trade. Friction occurs when two or more elements of culture come into
conflict, resulting in alteration or replacement of those elements. Innovation is the slight alteration of an
existing element of culture, such as a new style of dress or dance. Invention is the independent creation of a
new element of culture, such as a new technology, religion, or political form.

In the examples at the beginning of this section, building techniques and ideals move along with human
migration to new settings, where they must be altered to fit the materials and challenges of the new
environment. In colonial and neocolonial contexts, dominant groups may introduce the techniques and ideals
of their own homelands as “superior” even if they don’t work very well in the environments of colonial
conquest.

Some cultural inventions are so successful that they transform the whole way of life of a people. Consider the
information technologies that have reshaped American life since the 1970s, such as computers, the Internet,
and cell phones. These tools have changed the ways Americans communicate, work, learn, shop, navigate, and
entertain themselves. Diffusing through trade, these inventions have transformed cultures all over the world in
diverse ways. In many societies, modes of interacting through communication technologies come into conflict
with norms for interacting face-to-face, creating friction between the two realms. Where the movements,
behavior, and social relationships of young women are tightly controlled, for instance, mobile phones allow
women to secretly make new friends, explore new topics of conversation, and engage in behavior their elders
might not sanction.

Sometimes the forces of innovation and invention catch on, and sometimes they don’t. In the 1970s, Ralph
Hasty, a disc jockey from southwest Missouri, moved to Northern California, where he lived and worked for
many years. There, he learned about a new technology for building houses in the form of geodesic domes,
structures comprising intersecting polygons assembled from prefabricated kits. In late 1980s, he returned to
live in southwest Missouri, bringing with him this enthusiasm for geodesic construction. He ordered a kit and
built a geodesic dome house on a piece of rural land, intending to sell the house and use the profits to build
more of these geodesic wonders. Well, things did not exactly go to plan. The locals apparently found the house
far too weird to suit their notion of home. From the outside, the dome looked like some sort of futuristic
greenhouse or zoo habitat. On the inside, conventional furniture did not fit in the oddly shaped rooms of the
dome. Once finished, the geodesic home sat on the market for a number of months, and eventually, he had to
sell it at a loss. It must be mentioned that Ralph Hasty, geodesic innovator, continued to live in a conventional
rectangular house for the rest of his life.

FIGURE 3.13 Ralph Hasty stands in front of the geodesic dome he built. Although providing all of the needs of a
secure and warm dwelling space, it was hard to find a buyer for this unconventional home. (credit: Jennifer Hasty,
Public Domain)

Access for free at openstax.org


3.6 • The Paradoxes of Culture 89

Paradox 2: Culture Is Bounded but Mobile


Because many elements of culture are shaped by environmental forces, trading opportunities, and local
histories of settlement, culture becomes associated with territory. But because of the mobility of people,
objects, and ideas, culture rarely stays within the boundaries of any society; rather, it wanders restlessly along
lines of travel, communication, conquest, and trade.

People move around a lot, and this is nothing new. On the popular British television series Time Team,
archaeological excavations all over the United Kingdom uncover artifacts from ancient times that were
produced in far-flung places such as Rome, Scandinavia, and the Middle East. In episode 4 of season 16 (2015),
the team excavated a town in Wales that was constructed by Romans during the time of Roman conquest.
There, archaeologists unearthed the foundations of Roman buildings along with a variety of Roman objects,
including a third-century Roman coin, a Roman tool for removing earwax, a twisted-wire bracelet, and a knife
handle decorated with gladiators. Other Time Team investigations have uncovered artifacts from travelers and
pilgrims to sacred religious sites. These objects have diffused to British cultures through conquest, trade, and
migration. As people move around, so do objects, technologies, practices, and ideas.

FIGURE 3.14 This fabric shop displays a number of colorful wax print patterns. Although wax print fabrics are now
associated with Africa, the wax print technique actually originated in Indonesia. (credit: “National Colors” by
Miranda Harple for Yenkassa.com/flickr, CC BY 2.0)

However, certain integrated sets of things, practices, and ideas do cluster in certain places. Take a look at the
cloth in Figure 3.15. This kind of cloth is quintessentially African. It’s called wax print, and indeed, clothing
made of wax-print cloth is very popular in many parts of Africa. Wax-print cloth is industrially produced cotton
cloth with intricate designs and bold colors. In most African countries, a vast selection of designs and brands of
wax prints can be found in any market. Rather than buying ready-made clothes in clothing shops, people more
often purchase cloth in the market and take it to a seamstress or tailor to be made into the garment of their
own choosing.

Many wax-print designs are symbolic, serving as a means of nonverbal communication for the people who
wear them. Some cloths are associated with proverbs, occasions, monuments, and famous people. In the West
African country of Ghana, many cloth designs are named using the vivid proverbs of the large Akan cultural
group. One popular design features a bird in flight, associated with the Akan proverb Sika wo antaban,
meaning “money takes flight.” Another elaborate motif is called Akyekyde? Akyi, or “the back of the tortoise,”
worn by wise people who move through life with slow intention. One design with long, corrugated stripes is
called sugarcane, which is said to mean “I love you like sugar.”
90 3 • Culture Concept Theory: Theories of Cultural Change

FIGURE 3.15 The various designs on these fabrics are understood to each have a special meaning. In the upper
left, is an example of the Sika wo antiban design, meaning “money takes flight.” (credit: Ninara/flickr, CC BY 2.0)

Though iconically associated with African dress, wax print actually originated in Indonesia, derived from local
techniques for making batik cloth. Batik is made using wax to draw designs on plain cotton cloth that is then
immersed in a dye bath. When the wax is melted off, the design remains against the background of color. When
the Dutch colonized Indonesia in the 1700s, Dutch merchants were impressed with the beauty of local batik
and sought to use their own methods of mass-produced block printing to imitate the vibrant colors and
elaborate designs of Indonesian cloth.

In the 1880s, Dutch and British merchants introduced their own mass-produced wax prints to people in their
African colonies, particularly along the west coast of Africa. Dutch wax cloth was enthusiastically embraced by
Africans, who began to infuse certain patterns with social meanings. With independence in the mid-20th
century, many African countries developed their own wax-print textile industries using designs developed by
local artists.

Exemplifying the cultural paradox of locality and mobility, wax-print cloth is culturally embedded in African
culture while carrying a complex history of global trade, appropriation, and colonial domination.

In the context of global power relations, the mobility of culture poses questions about who has the right to
claim or use elements of culture diffused from elsewhere. As part of the process of cultural immersion and
participant observation, many cultural anthropologists adopt the dress, diet, gestures, and language of the
peoples they study while they are conducting fieldwork. Often, anthropologists bring their love of these
cultural elements back to their home societies and continue to use and practice them to show their
appreciation for the cultures they have studied. However, some people may find it unsettling to see a white
Euro-American anthropologist wearing an African wax-print dress—or a silk sari from India, or an ornately
woven lliclla cape from Peru. In your travels, have you ever purchased an item of clothing or jewelry worn by
local peoples? Is it appropriate to wear such items in your home society?

If someone is using cultural items as a way of honoring that culture, many people would think it’s perfectly
fine. If someone is wearing items from another culture as a form of humorous costume, such as a sports
mascot or Halloween costume, most people would find that offensive. An even more serious problem emerges
when a person uses or claims cultural elements from another society in order to make a profit. What if, for
instance, someone from the American fashion industry copied a wax print motif such as Sika wo antaban,
using the design for American clothing, housewares, or art? The elements of culture, both material and
nonmaterial, constitute the intellectual property of the people of that culture. Claiming or using the elements
of another culture inappropriately is called cultural appropriation.

Paradox 3: Culture Is Consensual but Contested


In any society, people interact using a set of assumptions about the sorts of behavior and speech considered

Access for free at openstax.org


3.6 • The Paradoxes of Culture 91

appropriate to certain people in certain situations. That is to say, culture is consensual; through their words
and actions, people agree to a certain way of doing things. As discussed earlier in this chapter, culture includes
conventionalized roles, behavioral norms, and shared ideas for framing situations.

For example, imagine that someone in the United States has just graduated from college and is looking for a
job. What should that person do? In the United States, it is common to spend time crafting an impressive
résumé, using a specific form of technical language that accentuates the quality of a person’s skills and
experiences while demonstrating their educational background. Instead of listing “worked as a camp
counselor,” someone might indicate that they “developed systems of cooperative leadership among youth in an
environmental awareness program.” A recent graduate would likely post this linguistic masterpiece to a job
search website such as Indeed.com.

For many people in China, such a strategy would seem very rudimentary and even grossly inadequate. Seeking
opportunities for education, employment, and business, people in China frequently rely on a cultural system
known as guanxi. Informed by Confucianism, guanxi refers to gifts and favors exchanged among people in
wide social networks based on mutual benefit. Guanxi is based on family ties but also includes relationships
formed in schools, in workplaces, and even among strangers who meet at parties or through mutual friends
(Yin 2017). While still in school, a student may be on the lookout for people who might be able provide access
to employment opportunities in the future. Using the practices of guanxi, the student would seek to establish
personalized links with such people in the hope that these links might prove advantageous in the future.

Say, for instance, a student hopes to get a job in solar technology after graduation. That student might seek out
professors whose teaching and research suggest connections in that industry. To establish relations of guanxi,
the student would not only take courses from that professor but also attempt to establish some sort of personal
rapport. This is typically done through strategic gift giving. In a particularly brutal winter, a student might knit
a sweater for the professor. An artistically inclined student might sketch a portrait of the professor and frame it
as a gift. Importantly, the gift must go slightly beyond the bounds of their professional relationship as professor
and student. Over time, the student might find ways to meet with the professor, further cementing the social
bond. After carefully cultivating this personalized relationship over months or years, the student might then
ask the professor to use industry connections to help them find a job.

What this means is that personal connections can be just as important as, if not more important than, the
language or qualifications of a person’s résumé. While Americans emphasize the importance of job-search
techniques, personal connections also play a role in securing employment in the American context,
particularly in highly paid, competitive industries such as software development and finance. In many
societies, people prefer to work with people they trust. Rather than hiring a random stranger, many prefer to
hire someone recommended by a trusted friend or business partner. In guanxi relationships, relations of trust
are established through the exchange of gifts and favors over time.

But what if the people who are hired in competitive industries are the ones who deployed their strategic social
connections and not necessarily the ones who are most skilled, talented, or otherwise best suited to the work?
What if the companies who are hired to complete infrastructure projects such as roads and bridges are not
necessarily the most competent or experienced ones but those who have given strategic gifts to government
officials? What if people use their guanxi networks to obtain special privileges, such as government licenses or
social services? Legal scholar Ling Li (2011) argues that some people use the cultural system of guanxi to
facilitate and rationalize bribery and other acts of corruption.

In 2012, the Chinese government launched an ambitious campaign against corruption among government
officials. More than 100,000 people have been investigated and charged with corruption, including many high-
ranking government officials, military officers, and senior executives of state-owned companies. Investigations
have revealed how powerful people use their extensive guanxi networks to secure deals, exert influence, and
extract goods and services. The campaign against corruption in China raises questions about the morality and
legality of guanxi practices.

Although guanxi is a widely accepted system for gaining access to goods, services, and opportunities, people
who don’t have elite connections may feel that this informal cultural system is unfair. For personal or ethical
reasons, some people may challenge or resist the practices of guanxi. Chinese journalist Lijia Zhang (2013)
92 3 • Culture Concept Theory: Theories of Cultural Change

describes how she was denied a promotion in her first job because she refused to give the expected guanxi
gifts to her boss. Zhang reports that most Chinese people complain about the widespread practices of
corruption but are forced to use their guanxi networks to get ahead in life.

Guanxi illustrates how culture can be generally taken for granted but also highly controversial. Many other
cultural norms are also widely accepted but challenged and resisted by certain groups who are disadvantaged
or limited by those norms. Gender roles are a good example, as are norms of sexuality and marriage.

Paradox 4: Culture Is Shared, but It Varies


The examples of guanxi and the geodesic dome both illustrate another paradox: how culture is widely yet
unevenly shared among members of a group. Different members of and groups in a society have different
perspectives on their shared culture—and different versions of that culture. Among elites, the use of Chinese
guanxi (or American “networking”) might seem to be a more personal and trustworthy process for making
things happen. But for people who lack access to elite networks, these cultural norms may seem to be an
exclusive and unfair tool of class oppression.

Returning to the notion of home, consider the many, many versions of home in your society. People in different
subgroups and regions live in structures of different shapes and sizes that are made of different materials. And
yet, the members of a culture do share a common set of assumptions about home. Home is where we live,
where we sleep, and most often where our family lives as well. Even with such diversity, people in a society
have a common image or ideal of home. On the West Coast of the United States, geodesic innovators sought to
expand the notion of home with a new shape and a new way of building. But in southwest Missouri, that
variation of home did not take root. Alas.

The four paradoxes all illustrate how culture operates as a force of stability in a society while also generating
forms of constant alteration, adaptation, and change. As culture is mobile, controversial, and variable, some
elements are always in the process of transformation even as other elements are maintained and reinforced.
Over time, people reinterpret their cultural norms and practices and sometimes even reject them altogether in
favor of some other way of thinking or doing things.

This paradoxical view of culture points to the dynamic tensions of people living in groups. Societies are
collectivities of individuals, families, regional groups, ethnic groups, socioeconomic classes, political groups,
and so on. Culture provides a way for people to live and work together while also allowing for the expression
and performance of distinctive differences. Rather than breaking down, culture responds to pressures for
change with adaptation to new conditions. The paradoxes that make culture seem impossible also make
culture flexible and durable. In an era that combines increasing polarization with an urgent need for
cooperative change, perhaps we need culture now more than ever.

MINI-FIELDWORK ACTIVITY

Romance over Time

Write down the answers to the following questions. What does a person in your culture do when they want to
become romantically involved with a particular someone? Are there common practices for this? What rules
guide this behavior, explicit or implied? What are the different roles involved? Are there symbols and rituals?
Is there some amount of disagreement in your culture about any of these activities?

Now, find a person in your culture who is much older than you, perhaps a person over 70 years old. Ask that
person to describe how people did the same things when they were your age. Ask the same set of questions,
and write down the answers.

How have romantic relations changed over time? What forces have shaped this change? What aspects have
remained the same? What explains the durability of some practices? Based on this trajectory of change, can
you predict how romantic relations will change in the future?

Access for free at openstax.org


3.6 • The Paradoxes of Culture 93

Suggested Readings
“Anthropological Theories: A Guide Prepared by Students for Students.” 2012. Department of Anthropology,
University of Alabama College of Arts & Sciences. https://anthropology.ua.edu/anthropological-theories/.

Bachmann-Medick, Doris, Jens Kugele, and Ansgar Nünning, eds. 2020. Futures of the Study of Culture:
Interdisciplinary Perspectives, Global Challenges. Boston: De Gruyter.

Geertz, Clifford. 1973. The Interpretation of Cultures: Selected Essays. New York: Basic Books.

Neumann, Birgit, and Ansgar Nünning, eds. 2012. Travelling Concepts for the Study of Culture. Boston: De
Gruyter.
94 3 • Key Terms

Key Terms
cultural appropriation claiming or using elements element of culture, such as a new technology,
of another culture in an inappropriate way. religion, or political form.
cultural ecology how humans develop culture as material culture objects made or used by humans,
an adaptation to various environments. such as buildings, tools, clothing, household
cultural evolutionism the study of the origins of items, and art.
human cultural forms and how those forms have norm the cultural expectations, including
changed over long periods of time. behaviors and attributes, that are associated with
cultural frames patterned, shared ways of a cultural role.
interpreting situations. ontological anthropology an approach that
cultural materialism an evolutionary approach explores how culture constructs our social and
that identifies technology and economic factors natural realities, what we consider real, and how
as fundamental aspects of culture, molding other we act on those assumptions. Reaching beyond
features of culture such as family life, religion, human realities, ontological anthropology also
and politics. attempts to include nonhuman perspectives,
cultural practices routine or habitual forms of relationships, and forms of communication.
behavior. ontology the study of the nature of existence.
cultural role a conventionalized position in a rite of passage a ritual that moves a person or
particular context or situation. group of people from one social category to
culture the whole way of life of a society, another, often more highly valued one. Examples
combining material objects, technologies, social of rites of passage include naming ceremonies,
relationships, everyday practices, deeply held initiations, weddings, and funerals.
values, and shared ideas. ritual repeated, patterned action conventionally
diffusion in an anthropological context, the spread associated with a particular meaning, often
of material objects, practices, and ideas among incorporating symbolic objects and actions.
cultures in complex relations of trade, migration, social structure the organizational framework for
and conquest. a particular realm of culture, such as the family,
friction occurs when two or more elements of the economy, or the political system. Social
culture come into conflict, resulting in alteration structures combine material culture with
or replacement of those elements. practices and ideas.
functionalism a form of analysis that focuses on structural functionalism a form of analysis that
the contemporary purposes of culture. describes how various aspects of culture fit
historical particularism an approach to cultural together and contribute to the integrated whole of
change that describes the combination of internal culture.
and external factors that shapes the unique structuralism the study of culture as a system of
historical trajectory of each culture. symbolic categories embedded in the myths,
ideology a model that depicts how a social realm religion, kinship, and other realms of a culture.
operates or should operate. An ideology identifies symbol an object, image, or gesture conventionally
the entities, roles, behaviors, relationships, and associated with a particular meaning.
processes in a particular realm as well as the technology specialized knowledge or skills
rationality behind the whole system. required to produce objects of material culture.
innovation the slight alteration of an existing unilineal evolution the idea that all cultures pass
element of culture, such as a new style of dress or through a single set of developmental stages.
dance. values cultural notions about what is good, true,
intellectual property material and nonmaterial correct, appropriate, or beautiful.
products of an individual or group that are worldview a very broad ideology that shapes how
protected by national and international laws and the members of a culture generally view the world
cannot be used for profit by others without and their place in it. Worldviews tend to span
attribution or compensation. several realms, including religion, economics,
invention the independent creation of a new and politics.

Access for free at openstax.org


3 • Summary 95

Summary
The discipline of anthropology is centered on the some examining change over time and others
concept of culture. What we make, what we do, and considering the functions of culture at one
what we think constitute the basic elements of particular point in time. While it is an incredibly
culture. These elements combine in aggregates such useful tool for understanding human social life, the
as symbols, rituals, and social structures. Since the concept of culture is riddled with paradox. Though
19th century, anthropologists have developed durable and integrated, culture is subject to
various modes of analysis for understanding culture, constant change, mobility, contest, and variability.

Critical Thinking Questions


1. Draw a floor plan for your ideal home. What those roles? Do you observe all of these norms, or
rooms would you have, and why? How would do you choose to ignore or resist some of them?
those rooms be organized? How is the imagined What happens when you publicly resist the
structure of this home shaped by an imagined norms of your culture?
lifestyle? What form of family or social relations 5. What rituals mark the passage of children into
are embedded in your house plan? What forms of adulthood in your culture? Identify symbolic
work? What notions about gender and age are objects and actions in those rituals. What norms
assumed? and values are expressed?
2. Describe your routine for getting ready in the 6. Which sports are popular in your culture or
morning. What aspects of this routine are region? Choose one. How might an
governed by biology, and what aspects are anthropologist use an evolutionary perspective
cultural? Ask a friend to describe their morning to analyze this sport? How might another
routine. Are there differences? Commonalities? anthropologist use a functionalist approach?
What norms and values shape these practices? 7. Do you believe in ghosts? Is this belief widely
3. List the colors of the rainbow. With a friend, shared among the people you know? How might
describe the symbolic meanings associated with the belief in ghosts shape cultural ideas about life
each color. Do you agree on these meanings? Do and death? Would ghost beliefs influence what
some colors have multiple meanings in your people do after a person dies?
culture? How do people use and interpret colors 8. Under what circumstances is it appropriate for a
with multiple meanings? person in one culture to adopt elements from
4. List the social roles you inhabit in your culture. another culture, such as dress, food, or speech?
What are the ideal behaviors associated with Is that cultural appropriation? Why or why not?

Bibliography
Bourdieu, Pierre. 1970. “The Berber House or the World Reversed.” Social Science Information 9 (2): 151–170.

Douglas, Mary. 1966. Purity and Danger: An Analysis of Concepts of Pollution and Taboo. London: Routledge &
Kegan Paul.

Ferguson, James. 2006. Global Shadows: Africa in the Neoliberal World Order. Durham, NC: Duke University
Press.

Geertz, Clifford. 1973. The Interpretation of Cultures: Selected Essays. New York: Basic Books.

Gennep, Arnold van. 1960. The Rites of Passage. Translated by Monika B. Vizedom and Gabrielle L. Caffe.
Chicago: University of Chicago Press.

Hallowell, A. Irving. 1992. The Ojibwa of Berens River, Manitoba: Ethnography into History. Edited by Jennifer
S. H. Brown. Fort Worth, TX: Harcourt Brace Jovanovich.

Kohn, Eduardo. 2013. How Forests Think: Toward an Anthropology beyond the Human. Berkeley: University of
California Press.

Li, Ling. 2011. “‘Performing’ Bribery in China: Guanxi-Practice, Corruption with a Human Face.” Journal of
Contemporary China 20 (68): 1–20. https://ssrn.com/abstract=1712390.
96 3 • Bibliography

Lyons, Harriet D. 2011. “Dame Mary Douglas.” In Fifty Key Anthropologists, edited by Robert J. Gordon, Andrew
P. Lyons, and Harriet D. Lyons, 46–51. New York: Routledge.

Peters-Golden, Holly. 2002. Culture Sketches: Case Studies in Anthropology. 3rd ed. Boston: McGraw-Hill.

Scalco, Patricia. 2019. “Weaving Value: Selling Carpets in the Liminal Space of Istanbul’s Grand Bazaar.”
Anthropology Today 35 (5): 7–10.

Tylor, Edward Burnett. 1873. Primitive Culture: Researches into the Development of Mythology, Philosophy,
Religion, Language, Art, and Custom. 2nd ed. 2 vols. London: John Murray.

Williamson, Kenneth. 2012. “Night Becomes Day: Carnival, Contested Spaces, and the Black Movement in
Bahia.” The Journal of Latin American and Caribbean Anthropology 17 (2): 257–278.

Yin, Xiangru. 2017. “An Analysis of Corruption in China: The Guanxi Network of Chinese High Level Officials
and Governors.” Master’s thesis, Clark University. International Development, Community, and
Environment (IDCE) (140). https://commons.clarku.edu/idce_masters_papers/140.

Zhang, Lijia. 2013. “Author: In China, ‘Everyone Is Guilty of Corruption.’” CNN. October 24, 2013.
http://edition.cnn.com/2013/10/23/opinion/china-corrution-lijia-zhang/.

Access for free at openstax.org


CHAPTER 4
Biological Evolution and Early Human
Evidence

Figure 4.1 The Grand Gallery of Evolution in the National Museum of Natural History in Paris, France displays
9,500 specimens of the estimated millions of species that currently live or once lived on the Earth in its collections.
In addition to educating the public about the mechanisms of evolution, the exhibitions in the museum honor the
scientists who helped contribute to our current understanding of the history of life on Earth. (credit: “Great Gallery
of Evolution” by Mustang Joe/flickr, Public Domain)

CHAPTER OUTLINE
4.1 What Is Biological Anthropology?
4.2 What’s in a Name? The Science of Taxonomy
4.3 It’s All in the Genes! The Foundation of Evolution
4.4 Evolution in Action: Past and Present
4.5 What Is a Primate?
4.6 Origin of and Classification of Primates
4.7 Our Ancient Past: The Earliest Hominins

INTRODUCTION This chapter applies evolutionary concepts to the understanding of human origins and
explains the biological variation seen in our ancestors across time. Chapters 4, Biological Evolution and Early
Human Evidence and Chapter 5, The Genus Homo and the Emergence of Us, represent a field of study that is
probably the most dynamic, controversial and highly debated subfield of anthropology. Perspectives and
opinions vary not only within the mindset of the general public but also amongst scientists and
98 4 • Biological Evolution and Early Human Evidence

anthropologists alike. As the human fossil puzzle begins to fill in with new discoveries, we find ourselves
gaining valuable insights into what makes us human and the ways in which we are a part of, not separate from,
the natural world. Despite our advances in the field, we also have to be prepared for the possibility we may end
up with more questions than answers! It is these very reasons that explain why so many of us find this such a
fascinating field and why so many of us take it so personally. It is after all a journey into the discovery of who
we are and where we came from; and that should be of interest to all of us as members of the Homo genus.

4.1 What Is Biological Anthropology?


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Identify the five subfields of biological anthropology.
• Explain how each of the subfields contributes to our understanding of human origins and evolution.
• Understand the historical context of the field of biological anthropology.

Looking to the Deep Past


Biological anthropology, also referred to as physical anthropology or evolutionary anthropology, is one of the
four major subfields of anthropology. While the other subfields focus on current and relatively recent human
cultures, biological anthropology looks to the deeper past, asking questions about what it means to be human
by exploring where humans came from as a species. Biological anthropology comprises numerous areas of
study: human biological variation, paleoanthropology (human and primate evolution), primatology (the study
of nonhuman primates), bioarchaeology (the study of bones found at archaeological sites), and genetic
anthropology (the application of molecular science to archaeological, historical, and linguistic evidence to
reveal the history of ancient human origins and migration). Each of these areas of study contributes something
to anthropologists’ understanding of current human physical characteristics and behaviors.

Exploring What It Means to Be Human


Studies of human biological variation evaluate the physical similarities and differences between human
populations across both time and space. Differences in morphology include features such as height, jawline,
eye sockets, and ear and nose shape and size. Biochemical differences account for variations in the sense of
smell, mutations in the CCR5 gene that offer resistance to HIV, and variations in skin pigmentation in response
to levels of exposure to ultraviolet rays from the sun.

FIGURE 4.2 These variations in modern human skin pigmentation are the result of evolutionary adaptations to
different levels of exposure to ultraviolet rays from the sun. (credit: “School Diversity Many Hands Held Together” by
Wonder woman0731/flickr, CC BY 2.0)

The study of human biological variation is closely linked to the original conception of biological anthropology,

Access for free at openstax.org


4.1 • What Is Biological Anthropology? 99

which was formalized in 1930 with the establishment of the American Association of Physical Anthropologists,
recently renamed the American Association of Biological Anthropologists. The change in name is an effort to
move away from the term physical anthropology, which has come to be associated with views promoting
scientific racism that no longer represent or align with views held by anthropologists today. In 1951, American
anthropologist Sherwood Washburn introduced a “new physical anthropology,” changing the focus from racial
typology and classification to the study of human evolution and the evolutionary process. This new focus
expanded anthropology as a field to include paleoanthropology and primatology

Paleoanthropology looks at the fossil evidence of humanity’s ancestors along with ancient material culture
such as tools and other human artifacts. The physical morphology (shape and size) of skulls and other
postcranial material (skeletal remains other than the skull) allow paleoanthropologists to form hypotheses
about important milestones in human evolution over time.

Primatology examines the behavioral and physical attributes of both living and fossil primates as well as their
relationships with their environments. Humans are primates who share a common ancestry with nonhuman
primates. By studying nonhuman primates, anthropologists can gain a better understanding of what it means
to be a primate and what it means to be human.

Genetic anthropology is used within several areas of biological anthropology. In this specialized area, DNA
testing is combined with archaeological, historical, and linguistic evidence to reveal the history of ancient
human migration or to track human disease.

Forensic anthropology is a subfield of biological anthropology that applies scientific methods to the analysis
of human remains for the purposes of identifying a victim and determining the possible cause of death. A
major difference between forensic anthropology and other types of biological anthropology is that forensic
anthropology is usually focused on crime scenes involving the death of an individual, whereas other types
primarily focus on understanding patterns and features that may appear in a group or an entire population.
Beginning in World War II, forensic anthropologists have been instrumental in helping identify victims of war
and disasters. They have played critical roles in identifying victims of the Thailand tsunami in 2004 and the
destruction of the World Trade Center on September 11, 2001. Today, most forensic anthropologists work in a
medical examiner’s office, assisting with autopsies and examinations of skeletal remains.

Bioarchaeology studies human remains in archaeological settings with a focus on what skeletal material can
reveal about the culture, diet, and presence of disease in a population. Bioarchaeologists are also interested in
the socioecological system of a population, which helps anthropologists better understand the roles of
environmental and ecological pressures and influences in shaping cultural identity, social inequity,
sustainability, and access to and use of resources. Based on the biological remains found at archaeological
sites, bioarchaeologists explore questions pertaining to social and funerary behavior, diet and nutrition,
health, and disease. Bioarchaeology offers a window into the connections among biology, society, and culture.
An example of what a bioarchaeologist might study is skeletal evidence of infant cranial boarding, which was
practiced by many cultures, including the ancient Maya, the Inca, and some Native North American groups.
The process involved binding a child’s head to a flat board in order to artificially deform the skull, possibly to
meet an aesthetic ideal or to signify social status. Bioarcheologists have found that variations in how the board
was attached to the skull provide important information about an individual’s social identity.
100 4 • Biological Evolution and Early Human Evidence

FIGURE 4.3 This elongated skull is from a member of the Nazca culture, which flourished in what is now Peru in
the years 100 BCE to 800 CE. It’s long, oval shape is the result of infant cranial bonding, the practice of deliberately
shaping the development of an infant’s skull by bonding it to stiff boards. (credit: “Nasca Peru Deformed Skull” by
VasenkaPhotography/flickr, CC BY 2.0)

PROFILES IN ANTHROPOLOGY

Ann Rosalie David


1946-

FIGURE 4.4 Professor Ann Rosalie David, Egyptologist and forensic and biological anthropologist at the University
of Manchester, UK. (credit: Professor David, Public Domain)

Personal History: Professor Ann Rosalie David was born in Cardiff, UK and earned a bachelor of arts degree in
ancient history from University College London in 1967 and a doctorate from the University of Liverpool in
1971. Her thesis was on ancient Egyptian temple rituals.

Access for free at openstax.org


4.2 • What’s in a Name? The Science of Taxonomy 101

Area of Anthropology: The focus of Professor’s David’s work has been biological anthropology and Egyptology.

Accomplishments In the Field: Professor David is a Director of the KNH Centre for Biological and Forensic
Studies in Egyptology at the University of Manchester. In this role, she established the Ancient Egyptian
Mummy Tissue Bank, one of the only such tissue banks in the world. She served as the keeper of Egyptology at
the Manchester Museum and has often worked in collaboration with Egypt’s Ministry of Health and Population
on public health projects. One such project involved the identification of antibodies against schistosomiasis, a
parasite spread by freshwater snails, in Egyptian mummies.

David was made an Officer of the Order of the British Empire (OBE) in 2003 for her work in Egyptology. David
has appeared in or consulted on several documentaries, including the television miniseries Private Lives of the
Pharaohs (2000) and Secrets of the Pharaohs (2001) and the documentary short Mummies: Secrets of the
Pharaohs (2007).

Importance of Her Work: Ann Rosalie David was the first woman in Britain to hold a professorship in
Egyptology. She was a pioneer in biomedical research, conducting research on disease, diet, and lifestyles in
ancient Egypt. In 2010, her work on ancient Egyptian mummies found evidence to suggest that cancer may be
a human-created disease, attributable in part to modern pollution and changes in lifestyle and diet (David and
Zimmerman 2010).

PODCAST
In this podcast (https://openstax.org/r/interview-mummies-withprof.annrosaliedavid), Professor Rosalie
discusses her work with ancient Egyptian mummies.

4.2 What’s in a Name? The Science of Taxonomy


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Describe the historical context of binomial nomenclature and scientific classification.
• Distinguish between the different categories of groups found in Linnaean classification.
• Explain the different definitions of species and how they are applied to different populations.

Defining the Science of Taxonomy


Taxonomy is defined as the classification and naming of things. Taxonomy organizes things into groups based
on predefined criteria. The criteria can be as simple as color or height or as complex as the presence or
absence of a trait, gene, or behavior. Taxonomy is a critical component of biological anthropology because it
helps anthropologists organize humans and their evolutionary ancestors both spatially (by location) and
temporally (through time).

Taxon refers to a specific subgroup, such as the genus. Taxa is the plural form of taxon, used to refer to all
groups. The classification system used for organizing living organisms was originally developed in the 18th
century by Swedish botanist Carolus Linnaeus. His system, which he called the Systema Naturae, uses a
structure known as binomial nomenclature. Binomial nomenclature assigns two Latin names to each
organism. The first is termed the genus name. The second is the specific or the trivial name, commonly called
the species name. In print, genus and species names are italicized. The first letter of the genus is capitalized,
while the species or trivial name is lowercase. For example, the scientific name for the house cat is Felis catus,
and the name for modern human beings is Homo sapiens. Linnaeus’s binomial nomenclature established a
shared scientific language that would become universal across countries and cultures, avoiding the confusion
caused by regional and colloquial names.

In addition to establishing a shared language, Linnaeus’s naming system groups organisms that share
common traits. For example, he grouped together animals with mammary glands into the category mammals.
Mammals were further broken down according to other traits. For example, mammals that have opposable
102 4 • Biological Evolution and Early Human Evidence

thumbs were grouped together as primates, and those without opposable thumbs were grouped as non-
primates. This is a hierarchical classification scheme, meaning that organisms are grouped into successive
levels from the broadest category of domain to the more specific level of species.

When Linnaeus first created his Systema Naturae, he built five hierarchical levels into his taxonomy: kingdom,
class, order, genus, and species. Humans are in the kingdom Animalia, the class Mammalia, the order
Primates, the genus Homo, and the species sapiens. Over time, many levels have been added to the Linnaean
system of classification, including domain, phylum, subclass, superorder, family, and tribe. The addition of
these taxon groups has enabled biological anthropologists to better understand the variations present in
various groups of organisms. However, biological anthropologists spend the majority of their time trying to
understand the species level.

FIGURE 4.5 This chart details the Linnaean hierarchical classification for the monarch butterfly. The broadest
category, “Life”, appears at the top of the chart, with classifications of increasing specificity at each level that
follows. “Species” is the most granular level. (attribution: Copyright Rice University/OpenStax, under CC By 4.0
license)

Defining a Species
While species is a word that most people are familiar with and comfortable using, just what determines a
species is incredibly difficult to define. At the most basic level, a species comprises a group of organisms with
shared characteristics that distinguish them from other groups. Most scientists distinguish a species based on
behavior, genetics, and/or morphology. Species definitions are the basis for scientific names. The common
name of a species, on the other hand, is usually based on general physical characteristics noted by a culture or
local population. Common names are also referred to as folk taxonomy or ethnotaxonomy (classifications
influenced by culture, etc.). There is a growing interest among anthropologists and the scientific community in
preserving Indigenous classifications of the natural world and connecting them with scientific classifications.

Decisions related to classification often involve tremendous taxonomical controversy, especially within the
field of biological anthropology. There are more than 20 distinct species definitions, or ways of categorizing or
distinguishing one type of organism from another. Below are the four most common definitions of a species.

Biological Species
The biological species definition states that a species is a group of interbreeding organisms that are

Access for free at openstax.org


4.3 • It’s All in the Genes! The Foundation of Evolution 103

reproductively isolated from other groups of organisms. Reproductive isolation means that members of a
species are not able to mate successfully with members outside their species. Gorillas, for example, cannot
successfully breed with Pan paniscus, the bonobo. The biological species definition uses the ability to
interbreed as its foundation because successful mating leads to gene flow, or the movement of genetic material
from one population to another.

Ecological Species
The ecological species definition emphasizes the role of natural selection in maintaining species boundaries.
This concept is based on the idea that gene flow is neither necessary nor sufficient to maintain species
boundaries. Instead, natural selection plays an important role in maintaining the boundaries between species.
In nature, species boundaries are often maintained even though there is a substantial amount of gene flow
between species. Gene flow between species generally occurs at places called hybrid zones, areas of overlap
where two species are known to successfully breed. A classic example of a hybrid zone occurs on the island of
Sulawesi in Southeast Asia, where Macaca maura (the moor macaque) and Macaca tonkeana (the Tonkean
macaque) are known to have successfully interbred for more than 150 years. Despite this, the integrity of the
two distinct species has been maintained.

Phylogenetic Species
The biological species definition is based on breeding behavior, specifically whether species are capable of
mating with one another. This foundation is problematic when trying to identify species over time. It is hard to
know whether two fossil specimens were capable of interbreeding. It is also difficult in the fossil record to
distinguish between interspecific variation (differences between members of two different species) and
intraspecific variation (variation within a species). Imagine finding the bones of two individuals, one five feet
tall and the other six feet four inches. Identifying whether these individuals were members of two different
species (interspecific variation) or representative of the normal variation within a given species would be
extremely challenging.

These problems are addressed by the phylogenetic species definition. The phylogenetic species definition
states that a species can be determined by shared possession of one unique characteristic. For example,
imagine you found a group of fossil leg bones. In order to decide if they were from the same species, you would
need to determine if they had a trait in common that only these fossil leg bones possessed. If the bones all
possessed trait A and this trait was not found in any other species already identified, then you would have a
new species, and all of the fossil leg bones could be placed in that species.

Mate Recognition Species


The mate recognition species definition states that a species is a set of organisms that recognize one another
as potential mates. A classic example of a group of species that can be distinguished using this definition is
American crickets. Within a single habitat in the United States, there might be over 30 different species of
crickets. Each species of cricket is known to produce a distinct song. Despite all these different species living
side by side, the female cricket of each species will only mate with a male after hearing the male sing her
species-specific song. The song, and the female recognition of it, constitutes a mate recognition system. This is
analogous to the biological species definition in that the song acts as a reproductive isolating mechanism.

4.3 It’s All in the Genes! The Foundation of Evolution


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Define alleles, genes, phenotypes, and genotypes.
• Distinguish the process of mitosis from the process of meiosis.
• Explain how Mendel’s laws of heredity affect human variation.
• Explain how the multitude of evolutionary forces contribute to variation in the human condition.

The Units of Life


Cells are the basic units of life in all organisms. They are the smallest entities that are capable of self-
reproduction. There are two main types of cells: prokaryotic and eukaryotic cells, named for the types of
104 4 • Biological Evolution and Early Human Evidence

organisms in which they occur. Prokaryotes are single-celled organisms, such as bacteria and archaea.
Eukaryotes are more complex, multicellular organisms, such as plants and animals (including humans). One
of the most important components of eukaryotic cells is the enclosed nucleus at the center of the cell;
prokaryotic cells do not have this nucleus. The nucleus of a eukaryotic cell houses all of the genetic material, or
DNA (deoxyribonucleic acid), that controls cellular function. Normally, the DNA forms a long string within the
nucleus.

There are two main types of eukaryotic cells: somatic cells and sex cells (also known as gametes). The somatic
cells make up the structural components of a body, such as the tissues, muscles, and organs. The sex cells are
specifically involved in reproduction. The function of the sex cells is to unite with a sex cell from another
individual to form a fertilized egg, also known as a zygote. In animals, there are two types of sex cells: ova, or
eggs, and sperm.

Cell division is the process that results in the production of new cells. However, sex and somatic cells divide
differently. The cellular division of somatic cells is known as mitosis, while the cellular division of sex cells is
known as meiosis. Mitosis of somatic cells is sometimes referred to as simple cell division because the parent
cell divides once to produce two daughter cells that are genetically identical to each other and identical to the
original parent cell. During mitosis, the DNA genetic material forms structures known as chromosomes. Each
daughter cell inherits an exact copy of all 46 chromosomes found within the parent cell.

FIGURE 4.6 In somatic cell division, also known as mitosis, the parent cell divides to produce two daughter cells
that are genetically identical to each other and to the parent cell. (credit: “Major events in mitosis”by Mysid/
Wikimedia Commons, Public Domain)

Meiosis, or sex cell division, is more complicated. This type of cellular division only occurs in the testes of
males and the ovaries of females. Instead of just one division, meiosis results from two cellular divisions that
produce four daughter cells. In meiosis, the four daughter cells each receive half of the original genetic
material from the parental cell. Thus, each daughter cell only has 23 chromosomes.

It is on the chromosomes that genes are housed. Genes are the fundamental unit of heredity. They are best
understood as the sequence or ordering of the DNA material that is housed in the nucleus. The genotype is the
genetic material found within an organism's cells and it is the expression of these genes that will produce the
phenotype or observable trait. Sometimes, the sequencing of the DNA material produces a variation of a gene,
known as an allele. An allele is defined as a similar but slightly different form of the same gene that can
activate the expression of a specific trait.

Gregor Mendel and the Laws of Heredity


The true nature of inheritance was not really understood until the beginning of the 20th century, when the
19th-century work of Gregor Mendel, a Catholic priest from Slovakia, was rediscovered. While in college,
Gregor Mendel was introduced to cell theory, which states that all organisms are composed of cells and that
cells are the fundamental unit of all living things. Cell theory raised many questions in Mendel’s mind,
including whether both parents contribute equally to the cells in their offspring. In 1854, Mendel began a
series of experiments with pea plants to help resolve this question and better understand how traits are
inherited from generation to generation.

Access for free at openstax.org


4.3 • It’s All in the Genes! The Foundation of Evolution 105

FIGURE 4.7 Gregor Mendel was a Catholic priest whose experiments with selective breeding of pea plants
established many of the rules of heredity. (credit: “Gregor Mendel Monk” by William Bateson, Mendel’s Principles of
Heredity: A Defence/Wikimedia Commons, Public Domain)

The first stage of Mendel’s experiments was identifying plants that breed true, meaning that each parent only
produces one kind of offspring when self-crossed. A self-cross is essentially a self-mating; some plants, such as
peas, have both male and female parts and can self-fertilize. Not all self-crosses are the same as the parent
plant, however. For example, self-crossed pea plants that have yellow pods sometimes produce offspring with
yellow pods and sometimes produce offspring with green pods. Mendel continued to selectively breed only
those pea plants that produced offspring that were the same as the parents. He called them purebreds and
referred to them as the P1 generation. It took him more than two years to establish plants that always bred
true.

Then Mendel selected seven traits of his pea plants that each had two distinct phenotypes, or observable
expressions of the trait. For example, seed shape can be either round or wrinkled, while pod color can be
either yellow or green. Over the next eight years, Mendel studied the mating and resulting traits of more than
28,000 plants. Mendel’s first round of experiments used his purebred pea plants to create what is known as a
monohybrid cross. A monohybrid cross is a mating between two purebred individuals who differ in a single
characteristic. In Mendel’s monohybrid crosses, the parent pea plants differed from one another in terms of
whether the pods of the parental pea plants were yellow or green or whether the seeds of the parental pea
plant were wrinkled or round.
106 4 • Biological Evolution and Early Human Evidence

FIGURE 4.8 Mendel identified a number of distinct characteristics observable in the seeds, flowers, pods, and
stems of pea plants. He used these observable traits as the basis for his breeding experiments, taking note of which
traits were dominant and which unexpressed (or recessive) in offspring. (credit: “Mendel Genetics” by LadyofHats/
Wikimedia Commons, Public Domain)

In his first monohybrid crosses, Mendel mated a purebred yellow pea plant with a purebred green pea plant.
He found that all the offspring resulting from this monohybrid cross were yellow, even though when the green
peas self-crossed, all their offspring were green. In other words, all the hybrid offspring were yellow in color. A
hybrid plant is one in whose parents differ in a term of a specific characteristic, such as pod color or seed
shape. The trait that was expressed (yellow) Mendel referred to as dominant, and the trait that disappeared
(green) he referred to as recessive. Mendel’s next set of experiments involved mating two hybrid plants—in
other words, those that resulted from the monohybrid cross. In these experiments, he found that the recessive
traits reappeared in a ratio of three dominant to one recessive.

FIGURE 4.9 These diagrams are examples of Punnett squares, a simple method for predicting the observable
results of breeding experiments. In the top square, a purebred green plant (yy) is crossed with a hybrid yellow plant
(Yy). The four possibilities for offspring appear in the four interior squares of the diagram. In this case, half of the
offspring will be hybrid yellow and half will be purebred green. The bottom diagram shows the results of a cross
between a purebred green and a purebred yellow plant – in this case, all offspring are hybrid yellow. (credit:
Copyright Rice University/OpenStax, under CC BY 4.0 license)

Mendel’s experiments suggested two very important facts. First, Mendel noted that various expressions of a
trait (such as pea color) were controlled by discrete units that occur in pairs and that offspring inherited one

Access for free at openstax.org


4.3 • It’s All in the Genes! The Foundation of Evolution 107

unit of each pair from each parent. This observation became Mendel’s first law of inheritance, the law of
segregation, which states that the two alleles for each trait segregate, or separate, during the formation of
gametes (eggs and sperm) and that during the reproductive process, the alleles combine at random with other
alleles. Today, we know that the process of meiosis—division of sex cells—explains Mendel’s law of segregation.
Each of the seven traits identified by Mendel is controlled by a pair of genes in the plant, one on each
chromosome. During the reproductive cycle, the chromosomes separate from one another so that each gamete
has only one allele for each trait. During fertilization, the alleles combine, and the two-gene state is restored.

After Mendel established his first law of inheritance, he extended his studies to more complex situations. He
began performing experiments with two set of traits, using dihybrid crosses. A dihybrid cross is a cross
between individuals who differ with respect to two gene pairs—for example, a cross between a plant with a
round yellow pea and a plant with a wrinkled green pea. Because yellow and round are both dominant traits
and wrinkled and green are both recessive, all the offspring resulting from the first-generation mating were
100% yellow and round. The green color and the wrinkled pea shape had disappeared. However, these
recessive traits reappeared in a ratio of three dominant to one recessive when two round yellow individuals
from the first-generation dihybrid cross were mated. The green color and the wrinkled pea shape had not truly
disappeared. In the second generation of the dihybrid cross, Mendel found that 9/16 of the offspring were
round and yellow, 3/16 were wrinkled and yellow, 3/16 were round and green, and 1/16 were wrinkled and
green. The results of these dihybrid crosses indicate that the two characteristics—pea color and pea
shape—segregate independently. The expression of one trait is not influenced by the expression of the other
trait. This is known as the law of independent assortment, which is Mendel’s second law of inheritance. There
is nothing to dictate that round peas will be yellow or that wrinkled peas will be green. The alleles that code for
different traits sort independently of one another during sex cell division (meiosis).

Mendelian Inheritance in Humans


Mendel’s laws of inheritance also apply to humans. Indeed, the principles of segregation and independent
assortment account for the transmission of certain human traits. Human blood type is one of the most familiar
Mendelian traits. Blood type has three phenotypes—A, B, and O—based on three alleles of a single gene. If only
the A allele or both the A and O alleles are present, the phenotype is A. If only the B allele or both the B and O
alleles are present, the phenotype is B. If both A and B are present, the phenotype is AB. If neither A nor B is
present, the phenotype is O. Note that O is recessive to both A and B, while A and B are codominant.
Codominance means that instead of one allele masking the other, the products of both alleles are observed.
Additional examples of Mendelian traits, or those controlled by a single gene, include Huntington’s disease,
widow’s peak, cystic fibrosis, sickle cell anemia, Tay-Sachs disease, hemophilia, and red-green color
blindness. OMIM, or Online Mendelian Inheritance in Man (https://openstax.org/r/onlinemendelian), hosts an
online database of almost 5,000 Mendelian human traits.

It is important to note that the majority of human traits are not controlled by a single pair of genes. More often,
a single gene can have multiple effects. Even more commonly, multiple genes are needed to produce a single
effect. These are referred to as polygenic traits. Most human traits are polygenic, not Mendelian. A good way of
determining if a trait is polygenic is to assess whether the trait can be measured. Traits that can be measured,
such as height or weight, are polygenic. Also, traits that have a wide range or lots of variability and can be
affected by environmental factors are probably polygenic. The survival of a species depends on genetic
diversity and variation. If there is a reduction of a gene pool due to geographic isolation or other environmental
factors than a species is at risk of extinction.
108 4 • Biological Evolution and Early Human Evidence

4.4 Evolution in Action: Past and Present


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Identify the major contributors to evolutionary theory and their specific theoretical contributions and historical
context.
• Explain the theory of evolution and how it applies to the understanding of human origins.
• Identify the key differences between Linnaean classification and phylogenetics.
• Define key evolutionary processes such as genetic drift, allopatric speciation, etc.

Contemporary biological anthropologists utilize an evolutionary perspective. This means that the principles of
evolution are used to understand how and why living organisms, including people, thrive in almost every
environment on Earth. More specifically, natural selection is accepted as the guiding force that shapes why
living things are the way they are. Out of all the possible variations of beings competing for the same resources
on Earth, those that prospered were the ones better suited to their environments than all other competitors.
The principles of evolution and natural selection will be discussed in some detail in the next few sections, but
it is important to establish at this early point that this chapter relies on the foundational assumption that
natural forces are the only forces directing the development of life on Earth.

Early Evolutionists and the Fixity of Species


Evolution is defined as change in the allele frequency within a gene pool that can lead to changes in an
organism’s morphology (form and structure) over time. Evolution involves the processes of mutation, natural
selection, and speciation, which will be introduced in upcoming sections. Prior to the 19th century, the
prevailing idea in Western thought was that nature was fixed and static; it was made by a supreme being in the
form it currently appeared, and it did not change. Within this fixed natural system, living creatures were
arranged within a set order that was considered to have been decreed by God, known as the great chain of
being. This order featured God at the top, angels beneath God, and then humans. Below humans were various
types of animals, followed by plants and minerals. This hierarchy was significant both because it placed some
creatures above others and because it distinctly separated humans from the rest of the animal world.

During a period stretching from the 14th through the 18th centuries, some people began to question whether
the natural world was as static as it was traditionally perceived to be. The British scientist and architect Robert
Hooke is remembered as the first person in the Western world to claim not only that nature has changed over
time but also that evidence of these changes remain. He hypothesized that fossils are the remains of actual
plants and animals that were once alive. This conclusion was contrary to the previously accepted conclusion
that fossils were nothing more than stone images. Hooke also noted that many marine fossils were located far
away from any existing ocean, and he came to the then radical conclusion that Earth’s geography and physical
features had experienced dramatic changes.

The first person to propose a mechanism by which species could change was French naturalist Jean-Baptiste
Lamarck, best known for having developed the first theory of macroevolution, a hypothesis about how the
actual transformation from one species into another species could occur. Lamarck’s theory relied on the now
defunct idea of the inheritance of acquired characteristics.

Lamarck argued that the usefulness of a trait or organ could be ascertained based on its complexity or size. In
particular, he believed that the usefulness of an organ could be judged by its size and the usefulness of a trait
by its complexity. He speculated that organs and traits that help a creature to survive will become bigger and
more complex over time, while those that are of little use will become smaller and simpler and eventually
disappear. His classic example of this theory in action is the long neck of a giraffe. Lamarck speculated that as
giraffes stretched their necks to reach the leaves at the tops of trees, their necks would grow longer, and
furthermore, these longer necks would be inherited by the subsequent generations. This theory of the
inheritance of acquired characteristics is also known as Lamarckian inheritance. One of the interesting things
about Lamarck’s theory is that he believed that wishes, desires, wills, and needs were all sufficient to motivate
change. That is, wishing for or desiring a change in one’s physical characteristics could make that change
happen.

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 109

There are two primary problems with Lamarckian inheritance. First, desires, wishes, and needs do not change
physical characteristics without a deliberate change in behavior. Someone may wish for blue hair, but their
hair color will not change without dye. The second problem is that the inheritance of acquired traits is not
possible. If someone dyed their hair blue, their children would not inherit blue hair. Traits that are acquired
during a lifespan are not passed on to subsequent generations.

Just because Lamarck’s theory of macroevolution is not correct does not mean that it is insignificant. Lamarck
recognized the importance of interactions between organisms and their environments in the evolutionary
process and was the first to propose a mechanism by which evolutionary change from one species into another
could actually occur.

Georges Cuvier, another Frenchman and a leading scientist in the early 19th century, made numerous
contributions to evolutionary thinking. He is best known for his theory of catastrophism, which he developed
to explain the increasing number of fossils that were being found, some displaying impressions of creatures no
longer found anywhere on Earth. Catastrophism proposes that floods, earthquakes, and other natural
disasters—understood within the theory as acts of God—have been responsible for killing all the animals alive
in certain places at certain times. According to Cuvier, either new animals have been created or the areas had
been repopulated by animals from neighboring areas. To be consistent with emerging fossil evidence
indicating that organisms had become more complex over time, Cuvier proposed that new organisms with a
more modern appearance were the result of a more recent creation event. While scientists no longer adhere to
catastrophism as a viable theory, Cuvier’s idea of extinction continues to be an important component of
evolutionary thinking today.

Another major contributor to evolutionary thinking was Scottish geologist Charles Lyell, known as the father of
modern geology. He wrote a three-volume treatise, Principles of Geology (1830–1833), in which he argued that
contemporary geological processes were the same as those that occurred in the past. These processes, such as
wind and rain, produced the contemporary geological landscape. Mountains, lakes, and rivers were all created
by these geological processes, many of them slow moving. This theory has come to be known as the principle
of uniformitarianism. Lyell suggested that in order for such slow-acting forces to produce momentous change,
Earth must be much older than previously suspected. Prior to Lyell’s publication, the majority of natural
historians believed that the earth was less than 6,000 years old, a number arrived at through calculations
made based on the Old Testament. By altering the suspected age of the earth from several thousand years to
millions of years, Lyell changed the framework within which scientists viewed the geological past.

Charles Darwin’s Role in Changing Views of the Natural World


Charles Darwin introduced a new way of seeing the world that was both highly criticized and acclaimed in the
scientific community of his time. In spite of resistance by various segments of society, his theories of natural
selection became the foundation of biological science. New knowledge pertaining to genetics and molecular
science has strengthened Darwin’s theories rather than weakened them.

Darwin the Apprentice


When he was 17 years old, well before he gained a reputation as a naturalist, scholar, and scientist, Darwin was
studying to be a medical doctor at the University of Edinburgh. Like many young people, he began to question
his original choice of studies, and he decided to instead learn taxidermy under John Edmonstone. John
Edmonstone was born enslaved and grew up on a plantation owned by a Scottish politician in what is now
Guyana in South America. Charles Waterton, the son-in-law of the plantation owner and a renowned naturalist,
would visit the plantation often. He started inviting Edmonstone to accompany him on his frequent travels into
the rainforest. On his travels, Edmonstone gained considerable knowledge about the flora and fauna of South
America along with impressive taxidermy skills.

After gaining his freedom in 1817, John Edmonstone taught taxidermy at the University of Edinburgh, where
he served as a mentor to Darwin over a period of several months. It is believed that Darwin’s relationship with
Edmonstone may have influenced his abolitionist views, which were later strengthened by firsthand accounts
of slavery while Darwin was on his infamous voyage to the Galápagos Islands off the coast of Ecuador.
110 4 • Biological Evolution and Early Human Evidence

FIGURE 4.10 At the University of Edenborough, John Edmonstone taught Darwin how to preserve birds. This is an
example of Embernagra platensis, the great Pampa-finch, collected by Charles Darwin in Uruguay in May of 1833.
(credit: “Embernagra platensis platensis, Great Pampa-finch, skin. Syntype. [B 19600]” by Michelle McFarlane/
Museums Victoria, CC BY 4.0)

Darwin the Explorer and Scholar


Charles Darwin left the University of Edinburgh and decided to pursue theology at Christ’s College, Cambridge.
His studies there led to his appointment in 1831 as a naturalist on the HMS Beagle for a five-year scientific
expedition around the world. During this voyage, Darwin collected, dissected, and organized various
specimens, especially in the Galápagos Islands, a chain of islands off the western coast of South America. His
observations in the Galápagos marked a crucial point in his thinking on evolution. He noted that the fauna and
flora of the western coast of South America were similar to those he observed in the Galápagos, but still distinct
enough to be considered different species. More surprisingly, the animals of each of the various islands in the
Galápagos chain differed slightly from one another. Darwin observed 13 different types of finches throughout
13 different small islands. The birds on each island differed in the structure of their beaks, their body form,
and the color of their feathers. Each species was specifically adapted to the specific habitats on each of the
islands. Darwin used the techniques that Edmonstone taught him to preserve the Galápagos finches, which
became key pieces of evidence supporting Darwin’s theory of natural selection.

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 111

FIGURE 4.11 Charles Darwin is acknowledged as the father of the theory of natural selection. His work built upon
the ideas of many other thinkers. His great contribution was in synthesizing these ideas into a coherent theory
explaining the diversity of life on earth and the great changes in life over geological time. (credit: A. C. Seward,
Cambridge Philosophical Society, Cambridge University Press/Wikimedia Commons, Public Domain)

During his travels on the Beagle, Darwin had been thinking about artificial selection—the selective breeding of
animals to produce traits that humans find useful, commonly associated with the process of domestication.
Darwin understood that artificial selection provided important clues about the natural evolution of species.

While on board the HMS Beagle, Darwin read a book by English economist Thomas Robert Malthus titled An
Essay on the Principle of Population (1798). Darwin obtained two important points from this book. The first
was that human populations, if unrestrained, will grow exponentially. This means that they will double each
generation. The second point was that food resources increase much more slowly than population does.
Malthus noted that the growth of human populations is kept in check by a limit of food resources, which
creates a struggle for existence. The struggle for existence is not just about getting enough food but also about
survival. In other words, it is about an individual’s ability to both find enough food and not become another
organism’s food. This simple concept, the struggle for existence, provided Darwin with a mechanism for how
evolution could occur. Darwin realized that individuals with favorable characteristics for living in an
environment are the ones that will survive to the age at which they reproduce, while those with less favorable
variations will not. This mechanism for “selecting for” certain traits and features is known as the theory of
natural selection.

Darwin concluded from his observations that when a group of animals of the same species are geologically
separated, they develop into separate species. This evolutionary process is commonly referred to as allopatric
speciation (or geographic speciation) and is based on the principles that related species share a common
ancestor and that species change over time.

Darwin did not originate the idea of evolution. Many of the ideas used by Darwin in his theory of natural
selection were developed by other thinkers. Darwin was also not the only person thinking about natural
selection. Another British natural historian, Alfred Russel Wallace, developed the same idea at roughly the
same time, entirely independently of Darwin. Whereas Darwin developed his ideas based on his travels to the
Galápagos, Wallace’s thinking was influenced by his own travels through the Malay Archipelago between
112 4 • Biological Evolution and Early Human Evidence

Indochina and Australia. Wallace outlined his theory of evolution by natural selection in a letter written to
Darwin while he was in Malaysia. As Darwin had not yet published his own work, Wallace and Darwin jointly
presented papers introducing the theory of natural selection. In 1859, Darwin finally published his book On
the Origin of Species, some 20 years after his voyage on the HMS Beagle.

Understanding Darwin’s Theory of Natural Selection


The theory of natural selection has five main components:

1. All organisms are capable of producing offspring faster than the food supply increases.
2. All organisms show variation.
3. There is a fierce struggle for existence, and those with the most suitable variations are most likely to
survive and reproduce.
4. Variations, or traits, are passed on to offspring (inherited).
5. Small changes in every generation lead to major changes over long periods of time.

A popular but often-misunderstood concept related to natural selection is the term survival of the fittest.
Survival of the fittest does not necessarily mean that the biggest and fastest survive; instead, it refers to those
who are most evolutionarily fit. This means that an organism has traits that are sufficient for survival and will
be passed on to future generations. The term survival of the fittest was not even introduced by Darwin; rather,
it was first used by English philosopher, anthropologist, and sociologist Herbert Spencer, who promoted the
now discredited ideology of social Darwinism. Social Darwinism applied the concept of Darwin’s biological
evolution to human societies, proposing that human culture was progressing toward the “perfect human.”
Spencer’s writings became integrally related to the 19th-century rise of scientific racism and European
colonialism.

FIGURE 4.12 This peppered moth is well camouflaged on the trunk of this tree. A darker colored moth would more
easily be seen and eaten and would thus be less likely to pass on its genes to offspring. Natural selection relies upon
the ability of natural variations to increase an individual’s chances of reproduction. (credit: Ben Sale/Wikimedia
Commons, CC BY 2.0)

Examples of Darwin’s theory of natural selection can be found throughout the natural world. Perhaps one of
the best known is the color change observed in peppered moths in England during the 19th century. Before the
Industrial Revolution, peppered moths in England were a light grey color, well camouflaged on tree branches
and less likely to be eaten by birds. Occasionally, through the process of mutation, black moths would appear
in the population, but these were usually quickly eaten because they were more visible against light-colored
bark. When soot from coal factories began to cover the bark of the trees, the black moths became better
camouflaged and the white moths were now more visible. Consequently, the black moths were the ones to
survive to reproduce, while the white ones were eaten. In a few decades, all the peppered moths in the cities

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 113

were black. The process was termed industrial melanism. As coal usage decreased and the bark of the trees
once again became lighter in color, white moths again dominated the urban areas.

Examples of natural selection in modern times are numerous. Pesticide resistance in insects is a classic
example. Pesticide resistance refers to the decreasing susceptibility of a pest population to a pesticide that
previously was effective at controlling it. Pest species evolve pesticide resistance via natural selection, with the
most resistant individuals surviving to pass on their ability to resist the pesticide to their offspring. Another
good example is the rise of “superbugs,” bacteria that have become increasingly resistant to antibiotics.

The Processes of Evolution


Mutation is the creative force of evolution and represents the first stage of the evolutionary process. Mutation
is defined as an alteration in a genetic sequence that results in a variant form. For a mutation to have
evolutionary significance, it must occur in the sex cells (sperm and ova). This is because only genetic
information that is in the sex cells is passed on from generation to generation. Mutations in non-sex
chromosomes will not be passed on from one generation to the next. Whereas other evolutionary forces can
modify existing genetic material, only mutation can produce new genetic material. One of the most interesting
things about mutations is the fact that they are random. There is no way of predicting when a specific mutation
will occur; all scientists can do is estimate the probability of a mutation occurring. Mutations do not
necessarily appear when they are needed.

The conventional view is that mutations are harmful, but this is not always true. Some mutations are harmful,
some are advantageous, and some are neutral. Advantageous mutations lead to changes that improve an
individual’s survival and/or chances of reproduction. The mutation that confers resistance to insecticide in
mosquitos led to changes that improved their survival. Likewise, the mutation for black coloration in peppered
moths led to increased survival during the Industrial Revolution. Neutral mutations have no effect on survival
or reproduction. And some mutations are in fact quite harmful and do negatively affect certain individuals’
survival and reproduction.

Mutations generally occur spontaneously in response to conditions in the body or in the environment. The
exact cause of a mutation cannot usually be determined, and the rate of mutation is very difficult to determine.
This is because mutations that are neutral or do not lead to obvious changes often go unnoticed. The
probability of a mutation at any given gene is between 1 in 10,000 and 1 in 100,000. While the probability that
a specific point in an individual’s genetic material will have a mutation is clearly very low, the probability that
the totality of an individual’s genetic material will have at least one mutation is much higher. The point is that
while rare, mutation is also common. For example, although many mosquitoes have adapted to insecticides
through a mutation that confers some resistance to the chemicals, if the mutation had not already been
present in the population, the mosquitoes would have died out. The need for a specific mutation had no effect
on whether the mutation appeared or not.

There is currently a controversial pilot program in Florida aimed at dealing with mosquitoes against which
insecticide sprays have increasingly become ineffective. The first genetically modified mosquitoes were
released in the Florida Keys in May of 2021. The genetically altered mosquitoes produce female offspring that
die in the larval stage, preventing them from growing to adulthood, in which they can then bite and spread
disease. Genetic science currently has the power to use mutations to control or even wipe out an entire
species. Genetic engineering has the potential to benefit humanity, but it will undoubtedly also raise ethical
questions and controversy.
114 4 • Biological Evolution and Early Human Evidence

FIGURE 4.13 Genetically modified mosquitoes are currently being bred that will die in the larval stage, thus greatly
reducing the mosquito population. (attribution: Rice University, OpenStax, under CC BY 4.0 license)

Genetic Drift
Genetic drift is defined as the effect of random chance on a population, notably the way in which it determines
whether an individual survives and reproduces or dies. Imagine that you stick your hand into a bucket filled
with Halloween candy. What is the probability you will withdraw a Snickers bar? The composition of Halloween
candy in your bucket will be affected by the proportion of people handing out Snickers bars compared to other
candy. If each bucket of Halloween candy were a population, then one could say that genetic drift—random
chance—was affecting the composition of the candy in your Halloween bucket. An important point about
genetic drift is that it is directly and inversely related to population size. The smaller the population, the larger
the influence of genetic drift; the larger the population, the smaller the influence of genetic drift. In a large
population, say 100,000, removing a couple of individuals will have a truly miniscule effect on the population.
Note that in early human evolution, however, population sizes were small, so the effect of genetic drift may
have been substantial.

Gene Flow
Gene flow is another important evolutionary force, involving the exchange of genetic material between
populations and geographic regions. Without gene flow, there would be no diversity—and without diversity, a
species is at higher risk of extinction. Gene flow can be seen in the process of pollination, in which bees or
butterflies carry and transfer pollen from one area to another. Anytime a gene is introduced to a new
population where it did not exist before, that is gene flow.

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 115

FIGURE 4.14 The process of pollination is a good example of gene flow. In this case, bees and butterflies transfer
genetic material, in the form of pollen, from one flower to another. (credit: “Honey Bee on a Dandelion, Sandy,
Bedfordshire” by Orangeaurochs/flickr, CC BY 2.0)

Speciation
Speciation is the rise of a new species in response to an environmental change or pressure. Allopatric
speciation, mentioned previously, is the most common form of speciation event. During allopatric speciation, a
species diverges when two populations become isolated from one another and continue to evolve. This
isolation is created by geographic barriers such as mountains, rivers, or oceans. A good example of allopatric
speciation is the different species of squirrel found on the two sides of the Grand Canyon. Descended from a
common ancestor, the squirrels became reproductively isolated from one another by the Grand Canyon,
eventually resulting in different species.

FIGURE 4.15 An example of allopatric speciation is the different species of squirrels that inhabit the Grand
Canyon. The squirrel on the left is a Harris antelope squirrel and the one on the right is a white-tailed antelope
squirrel. They look similar but are different species. (credit: left, “Harris Antelope Squirrel” by Saguaro National
Park/flickr, CC BY 2.0; right, “White-Tailed Antelope Squirrel” by Renee Grayson/flickr, CC BY 2.0)

Sympatric speciation involves species that are descended from a common ancestor and remain in one
location without a geographic barrier. A good example is the East African cichlid fish, which experience
reproduction isolation due not to a physical barrier but to females’ selection of mates with certain coloration.
The amount of light that reaches different levels and depths of the lake impacts how colors in the males appear
to the females. The East African cichlid fish are also a good example of adaptive radiation. Adaptive radiation is
seen when one or more species give rise to many new species in a relatively short time. Research shows that an
explosion of about 250 very diverse species of cichlids in Lake Tanganyika occurred in less than 10 million
116 4 • Biological Evolution and Early Human Evidence

years (Takahashi and Koblmüller 2011). Other research suggests that the common ancestor was the result of a
hybrid swarm from two different locations, as seen in Figure 4.16. (Meier et al. 2017).

FIGURE 4.16 There are more than 250 different species of East African cichlid fish, all traceable to two common
ancestors. The process through which a great number of species arises from a common ancestor within a relatively
short period of time is known as adaptive radiation. (credit: “1471-2148-5-17-3” by Phylogeny Figures/flickr, CC BY
2.0)

In peripatric speciation, members of the same population are separated and over time evolve as separate
species. Ring speciation is considered by some to be a type of peripatric speciation. Ring speciation occurs
when several species coexist for a time in a region near one end of a geographic barrier. When part of the
population migrates away from the original population (or gene pool) to the other side of the barrier,
reproductive isolation results. Reproductive isolation is strongest for that part of the population that is farthest
away from the original population. When too much variation has occurred between two groups, they will no
longer interbreed, and as a result, speciation—the development of two separate species—can occur. While fairly
rare, ring speciation is believed to explain the different species of the California salamander genus Ensatina.

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 117

FIGURE 4.17 This map shows the range of different species of the California salamander genus Ensatina, believed
to have developed through the process of ring speciation. In ring speciation, reproductive isolation leads to the
development of new species from a common ancestor, due to separation caused by distance and/or a physical
barrier. (credit: Thomas J. Devitt, Stuart J. E. Baird, and Craig Moritz/Wikimedia Commons, CC BY 2.0)

Gradualism vs. Punctuated Evolution


Biological anthropologists are interested not only in how a species is best defined but also in how often and by
what means new species are developed. The traditional view of evolution assumes that morphological,
behavioral, and genetic changes occur gradually and accumulate in a single unbroken and unbranching line;
this view of evolution is known as gradualism. If this perspective is correct, scientists would expect to find
numerous fossils exhibiting evidence that they are slowly and gradually transitioning into new and distinct
species. However, while fossils are rare, fossils showing evidence of transitional forms are even rarer. While the
dearth of transitional fossils is often attributed to the incompleteness of the fossil record, it has caused some
biological anthropologists to question if evolution is truly gradual.

What can be observed in the fossil record are static populations that are interrupted by sudden bursts of
change. This phenomenon of long periods of stasis, or no change, followed by quick periods of change is
known as punctuated equilibrium. Instead of a gradual accumulation of small changes, punctuated
equilibrium suggests that rapid changes due to a variety of environmental factors, including climate change,
are characteristic of the formation of new species. The fossil data for a large number of organisms show just
this—long periods of stasis followed by rapid and massive change. The scarcity of intermediary forms in the
fossil record has led some to conclude that punctuated equilibrium is the dominant theory. However, the fact
that intermediary forms do exist suggests that gradualism is also an important factor in the evolution process.
One research study found that 30 to 35 percent of speciation events occurred as the result of a sudden event or
change, while the remainder showed evidence of gradualism (Phillips 2006). In both the gradual and
punctuated models, speciation takes the form of branches through time rather than a linear progression.
Evolution is neither linear nor progressive, but rather a branching process—a tree of life containing both areas
of divergence and points of a shared common ancestry.
118 4 • Biological Evolution and Early Human Evidence

The Tree of Life: Showing Evolutionary Relationships

FIGURE 4.18 This sketch made by Charles Darwin illustrates his attempts to think through the branches of
evolutionary relationships. (credit: Charles Darwin/Wikimedia Commons, Public Domain)

During Darwin’s time, evolutionary relationships had to be determined largely by structural morphologies and
physical characteristics. Molecular science had not yet been developed. The binomial nomenclature discussed
earlier not only allowed distinction between species but also provided clues to evolutionary relationships. For
example, which of the below species of butterfly would be the most distantly related?

• Danaus gilippus
• Danaus genutia
• Limenitis archippus
• Danaus plexippus
• Danaus petilia

The answer, of course, would be Limenitis archippus, the viceroy butterfly, which is a mimic of the monarch
butterfly (Danaus plexippus). The first part of the viceroy’s name, Limenitis, is the genus. The fact that it is
different from the others shows that it is more distantly related.

FIGURE 4.19 Species can sometimes be difficult to identify by physical characteristics alone. The two butterflies in
this image are examples of two different species, one a monarch and the other a viceroy butterfly. What differences
can you see? (credit: left, “Today’s Mass Extinction and Holocene-Anthropocene Thermal Maximum” by
khteWisconsin/flickr, Public Domain; right, “A Viceroy Butterfly” by Benny Mazur/flickr, CC BY 2.0)

It is important to note that the Linnaean classification system has limits. Sometimes, species can be difficult to
identify by physical characteristics alone. Species that exhibit mimicry and larval forms in different stages of

Access for free at openstax.org


4.4 • Evolution in Action: Past and Present 119

development can take on the appearance of other organisms, resulting in errors in classification. Can you tell
which of the butterflies in Figure 4.19 is the monarch? Close examination reveals that the markings on the
wings are a bit different. The monarch is on the left, and the monarch mimic, the viceroy, is on the right.
Likewise, in Figure 4.20, you can see how it might be difficult to correctly classify barnacles, crabs, and limpets
based on physical appearances. One may be tempted to classify the barnacle and the limpet as being closely
related due to the conical shells that they share, when in actuality, the barnacle is more closely related to the
crab, as they are both crustaceans. The conical shells of the barnacle and the limpet are similar adaptations in
response to similar environmental pressures, not evidence that they are closely related or share a common
ancestor.

FIGURE 4.20 Classifying species based on physical similarities alone can lead to false conclusions. Although
barnacles and limpets look much more like one another than they do the crab on the left, barnacles are actually
more closely related to the crab. (credit: left, “DSC_5206” by Sally Wyatt/flickr, CC BY 2.0; top right, “Barnacles” by
Mo Riza/flickr, CC BY 2.0; bottom right, “Limpet Family at Sunny Cove” by Tim Green/flickr, CC BY 2.0)

Structural Morphologies as Evidence of Relationship


Structural similarities may be derived traits (homologous structures), inherited from a common ancestor, or
they may have developed independently (analogous structures). An example of a homologous structure is the
grasping hand found in both humans and chimpanzees, which suggests that humans and chimpanzees share a
common ancestor that also had a grasping hand. Analogous structures are seen in the wing of a butterfly and
the wing of a bat. While both wings serve a similar function, these two organisms likely developed their wings
independently and do not necessarily share a common ancestor. Identifying homologies is essential for
creating hierarchies of phylogenetic relationships because homology indicates that shared features are due to
common descent. However, homologies can be difficult to identify in nature, and they are easy to confuse with
analogous traits.
120 4 • Biological Evolution and Early Human Evidence

FIGURE 4.21 The structural similarities visible in these various species are homologous, meaning that the
similarities are the result of these animals sharing a common ancestor. (attribution: Rice University, OpenStax, under
CC BY 4.0 license)

Cladistics, or the use of cladograms, is a method of visually distinguishing between homologous ancestral and
derived characteristics. Ancestral characteristics are found in the common ancestor of the species being
classified, whereas derived characteristics are only found in the groups in question. An ancestral
characteristic that humans share with common ancestors is opposable thumbs. In contrast, a derived trait that
is only found in modern humans is the chin. By exclusively looking at derived characteristics, biological
anthropologists can develop a clearer understanding of the relationships between the groups being studied.

The Molecular Tree of Life and Phylogenetics


The emergence of genetic and molecular science has provided additional tools and lines of evidence to verify
evolutionary relationships. The phylogenetic tree is a model used by modern taxonomists to reveal the
complexity and diversity of life and its many branches. Phylogenetic trees show how species and other taxon
groups evolved from a series of common ancestors. They are based on both physical and genetic evidence.

FIGURE 4.22 Phylogenetic trees illustrate how old species are believed to be and their degree of relatedness to
one another. This particular tree pertains to primate species. (credit: Kosigrim/Wikimedia Commons, Public Domain)

Access for free at openstax.org


4.5 • What Is a Primate? 121

4.5 What Is a Primate?


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Define primate.
• Describe the relationship between primate behavior and environment.
• Identify and classify the key taxonomic groups of primates.

What Is a Primate?

FIGURE 4.23 Orangutans, the only great ape from Asia, are one of many living primate species. Others include
lemurs, monkeys, gibbons, and human beings. (credit: Dawn Armfield/Wikimedia Commons, Public Domain)

Primates—including human beings—are characterized by a number of distinct physical features that


distinguish them from other mammals. These include

• opposable thumbs and (in nonhuman primates) opposable big toes;


• the presence of five digits (fingers or toes) on the appendages;
• flat nails instead of curved claws;
• pads at the tips of the fingers made up of deposits of fat and nerves;
• reduced reliance on sense of smell and a relatively small snout;
• depth perception;
• binocular vision (being able to see one image with both eyes);
• a relatively slow reproductive rate;
• relatively large brain size; and
• postorbital bars (bony rings that completely surround the eyes).
122 4 • Biological Evolution and Early Human Evidence

FIGURE 4.24 The hands of this bonobo, including its opposable thumbs, look very similar to human hands.
Opposable thumbs or toes are a primate trait shared by no other group of mammals. (credit: “Bonobo Plankendaal”
by Marie van Dieren/flickr, CC BY 2.0)

The first four traits enhance dexterity and enable primates to use their hands and feet differently from other
mammals. Other traits on this list represent a shift in emphasis among the sense organs between primates and
other mammals. Primates are characterized by a greater emphasis on vision and a reduced reliance on smell
relative to other mammals.

Primate Behavioral Variation


Anthropologists regularly ask, “What makes us human?” Comparative studies of humans with nonhuman
primates help answer this question. Comparing the behavior of nonhuman primates and the behavior of
human beings helps anthropologists identify what culture is and develop operational definitions for it. Without
the comparative perspective provided by primatology, anthropologists would be missing an important piece of
the puzzle of what makes humans human. Without primatology, anthropologists would not be able to fully
understand humankind.

Studying nonhuman primates in their environment is key to understanding variations in behavior and can
shed light on humanity’s ancient past. Primatologists are studying the chimpanzees at Gombe National Park in
Tanzania, where they live in the rainforest. The behavior of chimpanzees that live in the tropical regions of
Africa is quite different from the behavior of chimpanzees that live in the savanna at Fongoli in Senegal, in
West Africa. Gombe chimps hunt red colobus monkeys without the use of tools, just catching them with their
hands, while the Fongoli chimpanzees hunt galagos (also known as bush babies) using sticks that they adapt
and used as spears (Pruetz, J.D, et al, 2015). The two environments also show differences in gender roles with
both males and females in the Fongoli savannah group involved in hunting while only male chimpanzees hunt
in the rainforests. Studying how these nonhuman primates both make and use tools is critical for
understanding how humans’ fossil ancestors may have used and constructed tools.

An important question that primatologists and biological anthropologists seek to answer is the question, do
nonhuman primates have culture? Whenever we see an exchange of ideas where one individual is involved in
teaching another and when that knowledge is passed on to others in a group is according to anthropologists, a

Access for free at openstax.org


4.5 • What Is a Primate? 123

form of culture. We see this happen in chimpanzee groups where older chimpanzees teach the young how to
use sticks to termite-fish, the process of extracting termites from a termite mound using a stick.

FIGURE 4.25 This chimpanzee lives in the Gombe National Park in Tanzania. Chimps living in Gombe’s rainforest
environment have developed a very different set of hunting techniques and tool use from their relatives living in the
grassy savannah. (credit: “Chimp Eden Sanctuary – Mimi” by Afrika Force/flickr, CC BY 2.0)

Explaining Primate Success


Why primates evolved as they did and how they filled and exploited the range of ecological niches they now fill
are questions that have not yet been adequately addressed. Over the last century, various hypotheses have
been raised to account for the evolution of primates and their unusual anatomical characteristics. These
theories include the arboreal theory, the visual predation hypothesis, and the angiosperm theory.

The arboreal theory proposes that primates evolved the traits they did as an adaptation to life in the trees.
Specifically, primates evolved thumbs and big toes that are perpendicular to the other digits to help them
grasp onto branches.

Matt Cartmill, a professor of anthropology at Boston University who spent his career trying to understand why
primates evolved the way they did, has complicated this theory. Cartmill recognized that forward-facing eyes
are characteristic not only of primates but also of predators such as cats and owls that prey on small animals.
Thus, forward-facing eyes, grasping hands and feet, and the presence of nails instead of claws may not have
arisen as adaptations to an arboreal environment. Rather, they may be adaptations that helped early primates
succeed as predators. According to the visual predation hypothesis, primate features are adaptations for
hunting insects and other small prey in the shrubby forest undergrowth and the lowest tiers of the forest
canopy.

The angiosperm theory states that the basic primate traits developed in coevolution with the rise of flowering
plants, also known as angiosperms. Flowering plants provide numerous resources, including nectar, seeds,
and fruits, and their appearance and diversification were accompanied by the appearance of ancestral forms
of major groups of modern birds and mammals. Some argue that visual predation is not common among
modern primates and that forward-facing eyes and grasping extremities may have arisen in response to the
need for fine visual and tactile discrimination in order to feed on small food items, such as fruits, berries, and
seeds, found among the branches and stems of flowering plants.

Primate Classification and Taxonomy


Scientists generally classify the order Primates into two suborders: Strepsirrhini (prosimians) and
Haplorrhini (tarsiers and anthropoids).

The Strepsirrhini or Prosimians


The Strepsirrhini are considered to be primitive primates that evolved much earlier than other primates. This
124 4 • Biological Evolution and Early Human Evidence

suborder includes lemurs and lorises. All the Strepsirrhini primates, or strepsirrhines, possess numerous
anatomical traits that distinguish them from the Haplorrhini primates, or haplorrhines. These include a
clawlike nail on the second toe, referred to as a grooming claw, and incisors in the lower jaw that are tightly
packed together and protrude from the mouth, forming what is called a toothcomb. There are seven families of
living strepsirrhines, and all of them are found in what anthropologists refer to as the Old World, which
consists of the continents of Africa, Asia, and Europe. Five groups of living strepsirrhines are found only on the
island of Madagascar off the coast of Africa. Two additional families are found in Africa and Asia.

FIGURE 4.26 The pygmy slow loris (Nycticebus pygmaeus) is an example of a Strepsirrhini primate. Pygmy slow
lorises be found in Vietnam, Laos, and a province of China. (credit: Lionel Mauritson/Wikimedia Commons, Public
Domain)

The Haplorrhini or Anthropoids


The Haplorrhini are broken down into two further infraorders, Simiiformes and Tarsiiformes, and the
Simiiformes are further divided into Platyrrhini and Catarrhini. The Platyrrhini, or platyrrhines, are
exclusively found in the New World (specifically Central and South America) and are colloquially referred to as
New World monkeys. Their name is derived from the rounded shape of their external nostrils, which open off
to the sides. New World monkeys are also distinguishable by their prehensile tails that serves as an extra limb
for extra support when moving in the trees. The Catarrhini, or catarrhines, are found throughout Africa and
Asia. They differ from the New World primates in that they possess narrow nostrils that face downward. The
Catarrhini contain two superfamilies, Cercopithecoidea and Hominoidea, and are exclusively Old World. The
Cercopithecoidea contain two main groups: cheek pouch monkeys (Cercopithecinae) and leaf-eating monkeys
(Colobinae). The most distinctive feature of the cercopithecoid primates is their molars, which exhibit two
parallel ridges. The most distinguishing feature of the hominoids is that they do not have tails and are largely
terrestrial, or ground-dwelling. Examples of Hominoidea include gibbons, chimpanzees, gorillas, orangutans,
and humans.

The Tarsier Puzzle


The tarsier, which belongs to the family Tarsiidae, has both prosimian and anthropoid characteristics, which
has made it difficult for scientists to classify. Tarsiers are currently classified within their own classification
under the haplorrhines. One of the characteristics that tarsiers share with other haplorrhines, (including
humans) is the inability to manufacture their own Vitamin C. They are the smallest known primate and are
nocturnal, with extremely large eyes that take up much of the space in their skull. Due to their size of the eyes,
the tarsier cannot rotate them; instead, it can rotate its head 360 degrees like an owl. Tarsiers are also the only
primate carnivore, eating largely flying insects and sometimes small animals like bats and lizards. Tarsiers do
not do well in captivity. They are extremely sensitive to noise and can become easily stressed. In fact, they can

Access for free at openstax.org


4.6 • Origin of and Classification of Primates 125

become so stressed that they die by suicide by banging their heads against tree trunks.

FIGURE 4.27 The Philippine tarsier (Carlito syrichta) is found only in the southern portion of the Philippine islands.
The tarsier has been challenging for scientists to classify, exhibiting both prosimian and anthropoid characteristics.
(credit: “8thApril2007 – ‘Tarsier’ Monkey” by Jacky W./flickr, CC BY 2.0)

4.6 Origin of and Classification of Primates


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Explain the concept of deep time.
• Define fossils and explain some dating methods used on fossils.
• Identify some of the key characteristics of early primate fossils, including their respective time periods.

Understanding Concepts of Time


Geologists divide deep history into time periods known as eras. Eras are generally based on the fossil life forms
observed. The oldest of the geological eras is the Eoarchean, which began approximately four billion years ago.
The majority of the fossil evidence that we have for primate evolution comes from the Cenozoic era—the
current geological era, dating from 65 million year ago (MYA) to the present. The Cenozoic era is divided into a
series of epochs. Each epoch is associated with specific forms of primates that evolved during that time period.

Fossils and Dating Methods


Biological anthropologists primarily, although not exclusively, study fossil artifacts. A fossil is any remainder
of a plant or animal that has been preserved in the earth. Upon the death of an organism, its body slowly
decomposes until all that remains are the teeth and the bones or a mere impression of the organism’s original
form. Under most conditions, teeth and bones and impressions eventually deteriorate, too. However,
occasionally conditions are favorable for preservation. Examples of favorable materials for fossil formation
include volcanic ash, limestone, and mineralized groundwater. Scientists do not have fossils of everything that
lived in the past, and in some cases, remains from only a few individuals of a species have been found. The
fossil record is very incomplete. Robert Martin, a curator at the Field Museum of Natural History in Chicago,
126 4 • Biological Evolution and Early Human Evidence

estimates that there have been more than 6,000 primate species, while the remains of only 3 percent have
been found. Fossils are very rare, but they are extremely informative about human biological evolution.

Making Sense of Fossils


An important part of understanding fossils is determining how old they might be and putting them in
chronological order. In order to use a primate fossil to reconstruct the evolutionary history of primates,
anthropologists must first be able to estimate approximately how old that specific fossil is. For some time,
relative dating methods were the only methods available for dating fossils. Relative dating calculates the
approximate age of a fossil in comparison to other fossil specimens. The last half century has seen important
advances in absolute dating, including techniques that have made possible the dating of the earliest phases of
primate evolution. Absolute dating calculates the actual biological age of a fossil in years within a range of
years.

Relative Dating Techniques


Stratigraphy is the best-known and most commonly used method of relative dating. Stratigraphy is based on
the observation that soil is deposited in successive layers, or strata. The oldest layers of soil (and any artifacts
or fossils within them) will appear beneath more recent layers of soil (and any artifacts or fossils within them).
In addition to using the location of layers of soil to date fossils deposited within these layers, biological
anthropologists also sometimes make use of other items consistently found in a specific layer of the soil. These
items are referred to as indicator artifacts because they help indicate the relative age of fossils and other
artifacts. The best indicator artifacts are those that have a wide geographic distribution, are presence for a
short period of geological time, and/or are from a species that underwent rapid evolutionary change. Different
indicator artifacts have been used to ascertain relative age in different areas of the world. In Africa, elephants,
pigs, and horses have been used to establish relative dates of different geological strata. The stratigraphy at
Olduvai Gorge in East Africa, for example, was established based on fossil pigs. The various species of pig in
successive strata are different and distinct, allowing paleoanthropologists to distinguish the strata based on
the pig species found within them. Once the stratigraphy of an area is established, the relative ages of two
different fossils in different sites can be determined by the associated indicator artifacts.

If a site has been disturbed, stratigraphy will not be a satisfactory way to determine relative age. In such a
situation, it may be possible to use absolute dating methods to estimate the age of fossils found together in a
disturbed site.

Absolute Dating Techniques


Many absolute dating methods are based on the rate of decay of a radioactive isotope. A radioactive isotope is a
chemical element that dissipates excess energy by spontaneously emitting radiation. These emissions happen
at known and stable rates. Once the rate of decay of a radioactive isotope is established, the age of a specimen
containing that isotope can be estimated within a range of possible error.

C-14
The best-known method for determining the absolute age of fossils is carbon-14 or 14C (pronounced “C-14”)
dating. All plants and animals contain the isotope carbon-14 (14C). Plants absorb 14C from the air, and animals
ingest plants containing the isotope. Because plants only absorb 14C when they are alive and animals only
consume plants when they are alive, scientists can determine how long ago an animal or plant died based on
the amount of 14C that remains in their cells. Carbon-14 has a known half-life of 5,730 years. This means that
approximately half of the original 14C in an organism will be eliminated in 5,730 years after its death. For
example, if an organism had an original 14C value of 100, then after 5,730 years, only 50 units of 14C would be
present.

Thermoluminescence
Another absolute dating technique that is frequently used by paleoanthropologists is thermoluminescence
dating. Thermoluminescence dating requires that either the fossils to be dated or the sediments that the fossils
are within have been exposed to a high-temperature event, such as a volcanic explosion. During such a high-
temperature event, all the radioactive elements within the material are released. Consequently, the amount of

Access for free at openstax.org


4.6 • Origin of and Classification of Primates 127

radioactive elements that have accumulated in the artifact since the time of the high-temperature event can be
used to calculate the artifact’s age.

Primates of the Paleocene Epoch


The Paleocene epoch began approximately 65 MYA and ended about 54 MYA. It is the most poorly understood
epoch of the Cenozoic era, as it is the time period with the fewest fossils to represent it. However, this epoch is
considered important to primate evolution because it offers the first unequivocal record of the earliest
primates. Evidence of the most primitive primate yet identified was found in the U.S. state of Montana, in a
geological deposit that was dated to the earliest part of the Paleocene. This creature is known as Purgatorius.
Purgatorius is similar to extinct and living primates – and distinct from other mammals – in the presence of an
elongated last lower molar and an enlarged upper central incisor (resulting in what one could think of as “Bugs
Bunny teeth”). These two characteristics, which are shared by all living primates today, suggest that
Purgatorius may be the common ancestor of later primates.

FIGURE 4.28 Purgatorius unio may be the common ancestor of all later primate. Remains of Purgatorius unio have
been found in deposits dated to be about 63 million years old. (attribution: Rice University, OpenStax, under CC BY
4.0 license)

Primates of the Eocene Epoch


The Eocene epoch, which began approximately 54 MYA and ended about 34 MYA, is marked by the
disappearance of Purgatorius and the first appearance of primates that more closely resemble modern-day
primates, especially in the fact that they possess postorbital bars composed entirely of bone. A postorbital bar
is a bony ring surrounding the entirety of the eye orbit. This contrasts with other mammals whose postorbital
bars are part bone and part cartilage. Some fossil specimens also possess a toothcomb and/or a grooming
claw, characteristics that are exclusively found in strepsirrhine primates today. Other anatomical
characteristics that are significant would be the ankle bones which researchers believe played a key role in the
evolutionary success of primates. The evolution of primates during the Eocene was tremendous. It has been
hypothesized that there were four times as many strepsirrhine primates during the Eocene than there are
living primates today. Fossil primates in Eocene deposits are common in North America and Europe and are
becoming known in Asia and Africa. However, there are currently no known fossil primates from the Eocene in
South America or Antarctica.

Primates of the Oligocene Epoch


The Oligocene epoch, which began approximately 34 MYA and ended about 22 MYA, marks the appearance of
the first fossil monkeys. The earliest unambiguous haplorrhine fossils were found at the Fayum, an
archaeological site about 60 miles from Cairo, Egypt, that today represents part of the Sahara. The Fayum
primates are divided into two main groups: Parapithecoidea and Propliopithecoidea. Based on their teeth,
these primates are believed to be the earliest New World and Old World monkeys, respectively. Teeth are
generally described according to a dental formula that indicates the number of each type of teeth in each
quadrant of the jaw. An organism with a 2.1.2.3 dental formula has two incisors, one canine, two premolars,
and three molars in each quadrant of their upper and lower jaws. Based on the presence of a third premolar, a
128 4 • Biological Evolution and Early Human Evidence

trait found in all New World monkeys, it is probable that Propliopithecus represents the earliest New World
monkeys, even though they first evolved in Africa. Likewise, it is probable the propliopithecoids represent the
earliest catarrhine primate, as they are the first fossil monkeys that possess a dental formula of 2.1.2.3 found
in catarrhine primates.

Miocene Apes
The Miocene epoch contains fossil evidence of some of the earliest apes such as Proconsul africanus africanus
which lived in Africa from 23 to 14 mya. The earliest Miocene ape, found in Africa, is Proconsul. Unlike
modern-day apes, the Proconsul lacked long, curved digits, suggesting that they were able to hang from
branches but more often moved about on all four of their limbs. Proconsul also lacked a tail, which is why they
are considered apes and not monkeys. Like all Old World monkeys and apes, including humans, their teeth
show a pattern of 2.1.2.3. Another well-known ape from the Miocene is Sivapithecus. Sivapithecus fossils are
very common throughout Asia, with a particularly large number having been found in Turkey. Like modern-
day humans, they exhibit very thick dental enamel, suggesting that these apes regularly ate very hard foods.
The most intriguing aspect of Sivapithecus morphology is that the skulls show a tremendous resemblance to
the living orangutan in features such as its tall nasal openings and high eye sockets.

FIGURE 4.29 Sivapithecus is one the earliest known ape species. Fossil remains exhibit the tall nasal openings and
high eye sockets currently visible in orangutans. (credit: “Sivapithecus indicus (Fossil Ape) (Dhok Pathan Formation,
Upper Miocene; Potwar Plateau, Pakistan)” by James St. John/flickr, CC BY 2.0)

While it is known that orangutans probably evolved from a Sivapithecus-type ape, there are no clear
candidates for the ancestors of modern African great apes. There have only been two fossils found that clearly
and unequivocally belong to the ancestors of modern African apes. Samburupithecus is a large late Miocene
ape found in northern Kenya. It is known to resemble modern African apes. It differs from other Miocene
fossils in having molar teeth that are elongated in a direction from the front of the mouth toward the back,
instead of from cheek to tongue. Another fossil from the late Miocene (9–10 MYA) that is sometimes identified
as an ancestor of modern African apes is Ouranopithecus, found in Greece, which has facial morphology that
links it to both African apes and humans.

Access for free at openstax.org


4.7 • Our Ancient Past: The Earliest Hominins 129

4.7 Our Ancient Past: The Earliest Hominins


LEARNING OUTCOMES:

By the end of this section, you will be able to:


• Compare and contrast some early hominin species.
• Identify some key adaptations and characteristics found in early hominins.
• Identify key adaptations and derived traits that emerged in changing environments.

Walking on Two Feet


The term hominin refers to all species considered to be in direct lineage to humans, which include the genera
Homo, Australopithecus, Paranthropus, and Ardipithecus. Hominids refers to all modern and extinct great
apes, which include humans, gorillas, chimpanzees, and orangutans and their ancestors. These terms have
been understood to represent different things over the years, but the definitions provided here are the most
current. While all hominins may differ in varying ways from one another, they all share one anatomical
behavioral complex: bipedal locomotion.

Scientists can hypothesize about how a creature moved by analyzing several aspects of its morphology.
Brachiators, animals that move by swinging from branch to branch, generally have long arms, while leapers,
animals that propel their bodies through the force of their lower limbs, have relatively long legs. Arboreal
primates have arms and legs of equal length. In bipedal locomotion, one leg is called the stance leg, and the
other is called the step leg. While the stance leg is on the ground, the step leg is off the ground and striding
forward. During normal walking, both feet are on the ground only about 25 percent of the time. As speed of
locomotion increases, the percentage of time that both feet are on the ground decreases. As a result, for most
of the time that bipedal organisms are moving, their body is balanced on only one of their legs (the stance leg).
To ensure that bipedal organisms do not fall over while balanced on their stance leg, they have undergone
many anatomical changes since the earliest hominin ancestors.

One of the most important anatomical changes that facilitate successful bipedalism is the angling of the femur
(upper leg bone) inward at what is referred to as a valgus angle, which positions the knees and feet under the
center of the pelvis. Bipedal hominins have also evolved spinal curves that make it possible for the hips to
balance the weight of the upper body. The evolution of the arch in the foot as well as the realignment of the big
toe so that it is parallel to the other toes is also instrumental in transmitting weight during the step phase of
bipedal locomotion.
130 4 • Biological Evolution and Early Human Evidence

FIGURE 4.30 In humans, the femur bones angle inward. This adaptation, known as the valgus angle, makes
bipedal locomotion (walking upright) more comfortable and more efficient. (attribution: Rice University, OpenStax,
under CC BY 4.0 license)

The most important evidence of early hominin bipedalism is provided by the work of English
paleoanthropologist Mary Leakey. In the 1980s, Mary Leakey discovered a 75-foot trail of footprints made by
three bipedal individuals who had crossed a thick bed of wet volcanic ash about 3.5 MYA. These footprints
were found in East Africa at the site of Laetoli. Based on the date and the location, it is probable that these
footprints were made by Australopithecus afarensis. Analysis of the Laetoli footprints indicates a modern
striding gait.

FIGURE 4.31 These replicas of the 3.6-million-year-old hominin footprints found in Tanzania by Mary Leakey are
on display at the National Museum of Nature and Science in Tokyo, Japan. (credit: “Australopithecus afarensis Fossil
Hominid Footprints (Pliocene, 3.6–3.7 Ma; Laetoli Area, Northern Tanzania, Eastern Africa)” by James St. John/
flickr, CC BY 2.0)

The evolution of hominin bipedalism required complex anatomical reorganization. For natural selection to
produce such a tremendous amount of change, the benefits of these changes must have been great. There have
been dozens of hypotheses for these changes, ranging from freeing hands to carry tools, food, or offspring to
increasing energy efficiency or thermoregulation (the ability to maintain the body’s temperature) by exposing

Access for free at openstax.org


4.7 • Our Ancient Past: The Earliest Hominins 131

more of the body’s surface. None of the hypotheses are testable, making it truly challenging to understand why
humanity’s ancestors made such a huge behavioral shift. The next sections explore some of the key discoveries
of early hominin fossils in which anthropologists see some of the earliest indications of the adaptation of
bipedalism in the human story.

Miocene Hominids
The first hominid fossils appear in the late Miocene, 10 to 5 MYA. Sometime between 7 MYA and 4 MYA,
hominids moved out of the trees and began to adapt more fully to a ground-based living niche. Unfortunately,
the fossil evidence from this time period is extremely sparse, but new finds continue to be discovered.

A complete cranium of Sahelanthropus tchadensis was found in 2002 by French paleoanthropologist Michel
Brunet and his team in Chad in West Africa. Sahelanthropus is a fossil ape that lived approximately 7 MYA and
is claimed by some researchers to be the last common ancestor of humans and chimpanzees. Genetic studies
indicate that humans and chimpanzees diverged from one another sometime between 5 MYA and 7 MYA, so
this species lived right at the time of the divergence. The cranial capacity is a mere 350 cubic centimeters (cc),
which is equivalent to a chimpanzee; the modern human cranial capacity is approximately 1,400 cc.
Sahelanthropus also has a very large brow ridge (the large bone above the eyes), and the location of the
foramen magnum, the opening at the base of the skull where the spinal column enters the skull, suggests that
its head was not held over its spine and thus it was not bipedal.

Orrorin tugenensis was found in Kenya in 2001 by geologist Martin Pickford of the Collège de France and
paleontologist Brigitte Senut of France’s National Museum of Natural History. Orrorin tugenensis was dated to
approximately 6 MYA. Orrorin was proposed to be a hominin due to anatomical traits that suggest bipedalism.
For example, the femoral head (the big, rounded ball at the top of the leg bone that connects the leg to the hip)
is much larger than in quadrupedal apes, suggesting the femur was being used to support the weight of the
upper body. The muscles attached to the femur also suggest bipedal movement. Another feature that suggests
that Orrorin is truly a hominin is the teeth, which exhibit thick dental enamel and small, square molars, much
like modern humans.

Pliocene Hominins
The Pliocene epoch extended from 5 MYA to 1.8 MYA. Fossils from the Pliocene show evidence of the evolution
of hominins that are clearly bipedal. They also show evidence of clear, albeit primitive, cultural behavior.
Climatically, the Pliocene was colder than the preceding Miocene, which resulted in changing sea levels and an
increase in ice at the poles, opening up some previously inaccessible areas. During this period, North and
South America became connected through the Isthmus of Panama, and a land bridge across the Bering Strait
appeared between Alaska and Siberia.

Ardipithecus ramidus
Ardipithecus ramidus was found in Ethiopia in 1992 by American paleoanthropologist Tim White and was
dated to about 4.4 MYA. This is the first discovered hominin species to be dated to the Pliocene era. Based on
the forward position of the foramen magnum, it can be concluded that Ardipithecus was bipedal. Also, the
upper arm bones are very small, suggesting that the arms were not used to support weight during locomotion.
Ardipithecus possesses numerous traits, such as thin dental enamel, evidence of a reduced canine, and an
opposable big toe. As a result of the latter trait, many believe that Ardipithecus was bipedal on the ground and
quadrupedal in the trees. This hypothesis is supported by the fact that the fossil bones were found in relatively
heavily forested environments. The reduced canine is a derived trait appearing even earlier than A. ramidus
and is not what we would typically see in African ape males who have large intimidating canines. Current
hypotheses suggest that over time smaller canines became dominant when there became less need to show
aggression along with a female preference for males with milder temperaments (Suwa, G., et al. 2021).
132 4 • Biological Evolution and Early Human Evidence

FIGURE 4.32 These skeletal remains have been identified as Ardipithecus, the first hominin species discovered
that has been dated to the Pliocene Era. (credit: Sailko/Wikimedia Commons, CC BY 3.0)

The Robust and Gracile Australopithecines


The next few sections will examine various australopithecine species that had diverse physical characteristics
related to morphology of the teeth and skull. Based on these characteristics, paleoanthropologists classified
these species into gracile and robust forms, as illustrated in Figure 4.33. Gracile species had a more
pronounced projection of the jaw (prognathism), less flared cheeks with no sagittal crest, and smaller teeth
and jaws. The sagittal crest in the robust australopithecines accommodated large temporalis jaw muscles for
chewing tough plant materials.

FIGURE 4.33 Australopithecine species are classified as either robust or gracile. A defining feature of the robust
species is the sagittal crest visible on the Paranthropus boisei skull on the left. Gracile species, such as A. afarensis,
on the right, display more pronounced projection of the face (prognathism). (credit: left, Rama/Wikimedia
Commons, Public Domain; right, “Australopithecus afarensis Fossil Hominid (Pliocene, Eastern Africa) 1” by James
St. John/flickr, CC BY 2.0)

Species considered to be the gracile include Australopithecus anamensis, A. afarensis, A. africanus, A. garhi,

Access for free at openstax.org


4.7 • Our Ancient Past: The Earliest Hominins 133

and A. sediba. The robust australopithecines (classified under the genus Paranthropus) include Paranthropus
robustus, P. boisei, and P. aethiopicus. The gracile species emerged around 4 MYA and disappeared 2 MYA,
while robust species continued to exist for another million years. The next sections will first take a look at
some of the gracile forms of australopithecine, followed by the robust forms.

Australopithecus africanus
Australopithecus africanus was the first australopithecine discovered, in 1924, and was described by
Australian anatomist and anthropologist Raymond Dart, who found the fossil in a box of fossils sent to him by
lime quarry workers at a site called Taung in South Africa. The most notable specimen in the box was a skull
from a child, which Dart had to chip away from the stone it was embedded in. It took Dart four years to
separate the teeth. The skull is now known as the Taung skull or Taung child. Dart argued that the Taung child
represents “an extinct race of apes intermediate between living anthropoids and man” (Wayman 2011). He
noted that the skull was long and narrow, not rounded as in modern humans, and its brain averaged a mere
422 cc, equivalent to a chimpanzee. However, the Taung child did not possess brow ridges, had circular orbits,
and had minimal prognathism as well as small canines and no diastema (space in the jaw for large canines to
be positioned when the mouth closes). These latter traits are all analogous to modern humans. Most
importantly, Dart noted that the forward position of the foramen magnum indicated that the skull was poised
on top of the vertebral column, suggesting bipedalism and an upright posture.

FIGURE 4.34 This partial skull is from a specimen known as the Taung child. The species, Australopithecus
africanus, displays traits that resemble modern humans in some ways but not others. (credit: Daderot/Wikimedia
Commons, Public Domain)

Australopithecus afarensis
In 1973, a good portion of a skeleton (about 40 percent) was found in the Afar region of Ethiopia by American
paleoanthropologist Donald Johanson. He called the skeleton Lucy, after a Beatles song. It was dated to around
3.75–2.8 MYA and was determined to be a member of the species Australopithecus afarensis. Like all fossils
recently discovered, Lucy was given an identification or accession number, KNM-AL-288. The KNM acronym
stands for the Kenya National Museum, where the fossil is housed, and AL stands for the Afar locality where
the fossil was found. Since then, more specimens of this species have been found in Kenya, Tanzania, and
Ethiopia, all in East Africa.
134 4 • Biological Evolution and Early Human Evidence

FIGURE 4.35 This child stands next to a recreated skeleton of A. afarensis. The long arms and long, curved fingers
and toes of A. afarensis are apparent. (credit: “Australopithecus afarensis Fossil Hominid (Lucy Skeleton) (Hadar
Formation, Pliocene, 3.2 Ma; Hadar Area, Afar Triangle, Northern Ethiopia, Eastern Africa) 2” by James St. John/
flickr, CC BY 2.0)

Australopithecus afarensis is dated from 3.9 to 2.9 MYA with an endocranial capacity of around 400 cc, which
is approximately the same as a common chimpanzee. There are two morphological features that provide
evidence that A. afarenis moved more like a great ape than a human. First, it had arms that were substantially
longer than modern humans’. Long arms are generally found in animals that hang from branches, suggesting
that A. afarensis also exhibited this behavior. Also, A. afarensis possesses finger and toe bones that are long
and curved, another characteristic of animals that hang from branches. However, there is one important
morphological feature of A. afarensis that suggests that this species may have moved somewhat like modern
humans. The shape of A. afarensis’s pelvis (hip bones) looks substantially more like a modern human’s than it
does an ape’s. Instead of the hip bones being long and narrow, they are short and wide. Most
paleoanthropologists believe that this change in pelvic shape indicates that A. afarensis moved like modern
humans do, on two legs. While A. afarensis may have locomoted bipedally, the morphological differences
between A. afarensis and modern humans suggest they did not move in exactly the same way. Current
consensus is that A. afarensis was both tree dwelling and bipedal. Other anatomical evidence of bipedalism
includes a more anterior position of the foramen magnum and the angle of the femoral head and neck.

Australopithecus garhi
Also found in Ethiopia, Australopithecus garhi is dated to approximately 2.5 MYA. Its cranial capacity is
slightly greater than A. afarensis, at 450 cc. Australopithecus garhi has incisors that are larger than those of
any of the known australopithecines or Homo. The function of the large incisors is not yet known. The most
exciting aspect of A. garhi is that it provides evidence of the earliest use of stone tools by a hominin.
Specifically, A. garhi fossils were found with fossil bones of ruminants, such as antelopes, that displayed
numerous cut marks. Cut marks are made on bones by the process of removing meat from the bones with
stone or metal tools. Based on this finding, biological anthropologists have hypothesized that A. garhi used
some type of stone tool for butchering.

Access for free at openstax.org


4.7 • Our Ancient Past: The Earliest Hominins 135

Australopithecus sediba
In 2008, the clavicle bone of Australopithecus sediba was discovered by Matthew Berger, the nine-year-old son
of American paleontologist Lee Berger, in Malapa, South Africa. Further excavation in a cave feature uncovered
two partial skeletons, one of an adult female and the other a younger juvenile. A. sediba is considered an
important species because it appears in the fossil record around the time of the first emergence of the genus
Homo around 2 mya. The classification of A. sediba was initially difficult to determine, due to its complex
overlapping features, which include humanlike spine, pelvis, hands, and teeth and a chimpanzee-like foot.
This combination of traits suggests both tree climbing and bipedal adaptations. After studying the
characteristics collectively, anthropologists classified A. sediba as a species of Australopithecus. It is
considered a direct ancestor of Homo erectus and Homo ergaster, which are discussed in Chapter 5, The
Genus Homo and the Emergence of Us . It is believed that A. sediba could be a descendent of A. africanus,
which suggests the species may be a dead end within the lineage to humans. Its classification and relationship
with the genus Homo will likely remain highly debated.

FIGURE 4.36 These bones are from Australopithecus sediba, which displays a humanlike spine and pelvis but a
chimpanzee-like foot. (credit: Phiston/Wikimedia Commons, CC BY 3.0)

Paranthropus robustus
Thirteen years after Raymond Dart’s discovery, South African paleontologist and medical doctor Robert Broom
discovered Paranthropus robustus at a site called Kromdraai in South Africa. The most obvious difference
between Dart’s and Broom’s respective fossils, A. africanus and P. robustus, is that the morphology of Broom’s
fossil is much larger. Its features include a sagittal crest and a flared zygomatic arch for the attachment of a
large temporalis muscle for chewing a diet reliant on hard nuts and seeds. This interpretation was further
supported by scanning electron microscopy (SEM), which was used to evaluate the markings etched into the
teeth. As the teeth increased in size the incisors and canines shrank, giving Paranthropus a flatter face with
less projection of the jaw. There are some who argue that depending on the environment and locale, some
Paranthropus may have been omnivores, with varied diets similar to those of H. ergaster. (Lee-Thorp,
Thackeray, and van der Merwe 2000).

Paranthropus boisei
Following in Broom’s footsteps, other scientists began searching for fossils in East Africa. Beginning in 1931,
Kenyan and British paleoanthropologist Louis Leakey and his wife, Mary Leakey, worked in what is known as
the Eastern Rift Valley, which is a 1,200-mile trough extending through Ethiopia, Kenya, and Tanzania. They
searched for almost 30 years before they found their first hominin fossil, Paranthropus boisei
(OH-5)—originally classified as Zinjanthropus boisei—in 1959. It is often referred to as the hyper-robust
hominin because of its mohawk of bone on the top of the skull. Other features include a low or absent forehead,
a flat face, large jaws, and large attachment sites over the entire skull for chewing muscles.
136 4 • Biological Evolution and Early Human Evidence

Paranthropus aethiopicus
We have little knowledge about Paranthropus aethiopicus (shown in Figure 4.37), which has been dated to
about 2.5 MYA and is referred to as the “black skull.” It is believed that this species falls somewhere between
the robust and gracile australopithecines, having characteristics of both. The species was discovered in
Ethiopia in 1967 by a French expedition team headed by Camille Arambourg and Yves Coppens.

FIGURE 4.37 Much remains to be learned about Paranthropus aethiopicus, which has characteristics of both the
robust and gracile australopithecines. (credit: “Paranthropus aethiopicus (Fossil Hominid) (Nachukui Formation,
Upper Pliocene, 2.5 Ma; Lomekwi, Lake Turkana Area, Kenya) 3” by James St. John/flickr, CC BY 2.0)

Landmarks and Questions


While the fossils discovered up to this point have provided a small window into the story of humanity’s past,
they have also simultaneously raised numerous questions. Questions related to phylogenetic relationships and
points of divergence are challenges for paleoanthropologists, who have only fragmentary fossil evidence to
build hypotheses around. Nevertheless, the discoveries that have been made represent important landmarks
in anthropologists’ understanding, providing clues that will lead to the next steps in the human journey.

MINI-FIELDWORK ACTIVITY

Pedestrian Survey

Conduct a pedestrian survey to try to locate fossils near where you live (trilobites in New York, ammonites in
Texas, shark teeth near riverbeds, arrowheads). Think about where you would most likely find a fossil and why.
Try to extract one without destroying the environment around it, which provides important context. Try to
figure out what kind of fossil it is by doing some Internet research. Why do you think that this fossil was
preserved? What information would make the search for fossils easier?

Access for free at openstax.org


4 • Key Terms 137

Key Terms
allele an alternative form of a gene that arises by derived characteristics physical traits that are
mutation and is found in the same place on a present in related organisms but absent from
chromosome, directly impacting the expression their last common ancestor. They are often
of a genetic trait or phenotype. associated with a speciation event.
allopatric speciation speciation that occurs when diastema a space or gap between the canines and
two populations of the same species become the other teeth that allows for the upper and lower
isolated from each other due to a change in the teeth to bite together.
environment, such as geographic isolation. ecological species definition a definition of
analogous structures anatomical similarities species that explains differences in form and
between two species that suggest not a common behavior as the result of adaptations to the
ancestor but rather similar environmental environment and natural selection.
adaptations. ethnotaxonomy the study of organism
ancestral characteristics homologous structures classifications and taxonomies developed and
or traits that may also be found in the common used largely by Indigenous peoples and other
ancestor of the species being classified. cultural groups.
angiosperm theory a hypothesis that suggests that evolution changes that appear in a species over
primate origins and typical primate time. Evolution is dependent on genetic variation
characteristics developed in response to the and natural selection to pass on beneficial traits
emergence of flowering plants. that will increase survival of the species.
arboreal theory a hypothesis that proposes that foramen magnum the opening at the base of the
primates evolved the traits they did as an skull where the spinal column and nerves enter to
adaptation to life in the trees. reach the brain. The position of the foramen
artificial selection the process of deliberately magnum can be used to determine if a species
breeding certain specimens of plants or animals was bipedal.
to encourage desired traits. forensic anthropology a branch of biological
binomial nomenclature the scientific naming anthropology in which scientific techniques are
system developed by Carolus Linnaeus that used to determine the sex, age, genetic
represents two parts of a taxonomic name. The population, or other relevant characteristics of
name is italicized, the genus is always capitalized, skeletal or biological materials related to matters
and the species is always lowercased. For of civil or criminal law.
example: Homo sapiens. fossils any remains, impression, or traces of living
bioarchaeology the study of bones and other things from a former geologic age.
biological materials found in archaeological gene flow alteration of the frequencies of alleles in
remains. a population that results from interbreeding with
biological species definition a definition of organisms from another population.
species as members of populations that actually genetic anthropology a branch of biological
or potentially interbreed in nature. anthropology that uses molecular science to
Catarrhini a subcategory of the primate infraorder explore questions concerning human origins,
Simiiformes that includes any primate early human migrations, and the appearance of
considered an Old World monkey, an ape, or in disease across time.
the lineage of humans. This classification genetic drift random changes in the frequencies of
features downward-facing nostrils and a 2.1.2.3 alleles in a gene pool.
dental formula. genotype a complete set of genetic material found
catastrophism the theory that changes in Earth’s in an organism.
fauna and flora were caused by supernatural gradualism the idea that species evolve slowly and
catastrophic forces rather than evolution. continuously over long periods of time.
Cercopithecoidea a superfamily of the primate great chain of being a concept detailing a
infraorder Simiiformes, subcategory Catarrhini, hierarchical structure of all matter and life.
that consists of Old World monkeys. Haplorrhini a suborder of primates that contains
cladistics the classification of organisms based on tarsiers, New World monkeys, Old World
branchings of descendent lineages from a monkeys, apes, and humans.
common ancestor hominid the group representing all modern and
138 4 • Key Terms

extinct great apes, including humans, the earliest New World monkeys, though they first
chimpanzees, gorillas, orangutans, and all their evolved in Africa.
immediate ancestors. phenotype the set of observable characteristics or
hominin the group representing modern humans, traits of an organism, such as color and structural
extinct human species, and all of humanity’s morphology.
immediate ancestors, including the genera Homo, phylogenetic species definition a definition of
Australopithecus, Paranthropus, and species based on individuals all possessing
Ardipithecus. specific derived traits.
Hominoidea a superfamily of the primate Platyrrhini a subcategory of the primate
infraorder Simiiformes, subcategory Catarrhini, infraorder Simiiformes that comprises New
that consists of gibbons, great apes, and World monkeys
humanlike primates, including Homo and related polygenic traits traits that are controlled by
fossil forms. multiple genes instead of just one.
homologous structures similar anatomical primatology the branch of biological anthropology
structures that appear in different species and dealing with the primates.
suggest a common ancestor. Proconsul a genus of ape from the early Miocene.
hybrid zones areas where two distinct species prognathism projection of the face, as seen in
mate and produce offspring many nonhuman primates and early hominins.
industrial melanism the prevalence of dark- Propliopithecoidea a superfamily of primates
colored varieties of animals (for example, from the early Oligocene that is related to Old
peppered moths) in industrial areas where they World monkeys and is believed to represent the
are better camouflaged against predators than earliest catarrhine primate.
paler forms. punctuated equilibrium a hypothesis holding that
inheritance of acquired characteristics the the evolution of species proceeds in a
disproved idea that an organism can pass on to its characteristic pattern of relative stability for long
offspring physical characteristics that it has periods of time interspersed with much shorter
acquired during its lifetime. periods during which many species become
interspecific variation the genetic variation seen extinct and new species emerge.
between two species. Purgatorius genus of the earliest primate or proto-
intraspecific variation the genetic variation seen primate.
within a species. reproductive isolation conditions that prevent
law of independent assortment a law of potentially interbreeding populations from
inheritance stating that different genes and their breeding.
alleles are inherited independently. socioecological system the interrelationship
law of segregation a law of inheritance stating that between the diversity of plants and animals,
when two alleles for a trait separate during the humans’ environments, and the diversity of
formation of new zygotes, these alleles will human culture and language.
combine at random with other alleles. species a class of individuals that have some
morphology the physical shape and structural common characteristics or qualities.
form of an organism or species. stratigraphy a branch of geology dealing with the
mutation a change in the structure of a gene that classification, nomenclature, correlation, and
results in a variant form that may be transmitted interpretation of stratified rocks.
to subsequent generations. Strepsirrhini a suborder of primates that includes
natural selection the process by which a species lemurs, lorises, and galagos (bush babies).
that is able to adapt and to pass on beneficial survival of the fittest the theory that the most
traits to its offspring ensures survival of the evolutionarily fit members of a species will pass
species; first formally introduced by Charles on their traits to later generations.
Darwin. sympatric speciation speciation without a
paleoanthropology the study of the origins and geographic barrier.
predecessors of the present human species based taxa the plural form of taxon, used to signify all
fossils and other remains. taxonomic groups.
Parapithecoidea a superfamily of primates from taxon a specific group or subgroup of organisms.
the early Oligocene that is believed to represent taxonomy the science or technique of naming and

Access for free at openstax.org


4 • Summary 139

classifying life. visual predation hypothesis a hypothesis that


uniformitarianism the concept that Earth’s explains the origins of unique primate traits as
surface was shaped in the past by slow-moving adaptations for preying on insects and small
geological processes. animals.

Summary
Biological anthropology strives to understand how not affect the probability that an individual will have
humans interact and behave in the present, how another trait.
humans evolved biologically, and how humanity’s
Carolus Linnaeus is best known for creating the
ancient ancestors lived in diverse climates and
classification system that taxonomists use today,
environments. The anthropological approach to
which is based on physical similarities and
exploring these questions is grounded in
differences. Phylogenetics is a hypothesis about how
evolutionary theory. Charles Darwin was one of the
species are related to one another and to a common
first to propose a mechanism by which evolution
ancestor. Today, biological anthropologists apply
occurred, which he called natural selection. Natural
taxonomies and phylogenies to the current
selection is based on the premise that those with
nonhuman primate and hominin fossil record. It is
more favorable characteristics survive and
in the Miocene that the first fossil apes, such as
reproduce at greater rates than those without them.
Proconsul, are seen. The first evidence of hominin-
Natural selection depends on the evolutionary
like fossils appears by the end of the Miocene. A
processes of mutation, speciation, gene flow, and
large number of morphological changes observed in
genetic drift.
early hominins suggest considerable environmental
Darwin’s theory did not address how these favorable and climatic change. During the Pliocene epoch,
characteristics could be inherited. Gregor Mendel’s extending from 5 to 1.8 MYA, the evolution of
experiments on peas addressed this very question. hominins that were clearly bipedal is evident in the
Mendel’s work resulted in two very important fossil record, as is evidence of cultures that used
observations. He observed that the two alleles for stone tools. The path is now ready for the next group
each trait separate during the formation of the sex in humanity’s evolutionary history to enter the
cells and that the probability of having one trait does scene.

Critical Thinking Questions


1. How do anthropologists define being human? fossils?
2. What are some of the key differences between 6. What is the difference between a prosimian and
Linnaean classification and phylogenetics? an anthropoid?
3. In what ways does the COVID-19 virus exhibit 7. What are some of the main differences between
natural selection? Old World and New World monkeys?
4. How do anthropologists define a primate? 8. What are some of the key characteristics seen in
5. What is the difference between absolute and early hominins, and what environmental forces
relative dating methods? What methods are may have contributed to those changes?
commonly used when working with hominin

Bibliography
Aiello, Leslie C. 1986. “The Relationships of the Tarsiiformes: A Review of the Case for the Haplorhini.” In Major
Topics in Primate and Human Evolution, edited by Bernard Wood, Lawrence Martin, and Peter Andrews,
47–65. Cambridge: Cambridge University Press.

Blaxland, Beth. 2020. “Hominid and Hominin—What’s the Difference?” Australian Museum. February 10,
2020. https://australian.museum/learn/science/human-evolution/hominid-and-hominin-whats-the-
difference/.

Crompton, Robin. 1989. “Mechanisms for Speciation in Galago and Tarsius.” Human Evolution 4 (2): 105–116.

Dasgupta, Shreya. 2019. “Super Variable California Salamander Is ‘an Evolutionist’s Dream.’” Mongabay. March
18, 2019. https://news.mongabay.com/2019/03/super-variable-california-salamander-is-an-evolutionists-
140 4 • Bibliography

dream/.

David, A. Rosalie, and Michael R. Zimmerman. 2010. “Cancer: An Old Disease, a New Disease or Something in
Between?” Nature Reviews Cancer 10:728–733. https://doi.org/10.1038/nrc2914.

Dunbar, Robin I. M. 1988. Primate Social Systems. Ithaca, NY: Cornell University Press.

Fleagle, John G. 1999. Primate Adaptation and Evolution. 2nd ed. San Diego: Academic Press.

Fuentes, Agustín. 2019. Biological Anthropology: Concepts and Connections. 3rd ed. New York: McGraw-Hill
Education.

Groves, Colin. 2001. Primate Taxonomy. Washington, DC: Smithsonian Institution Press.

Hill, W. C. Osman. 1955. Primates: Comparative Anatomy and Taxonomy. Vol. 2, Haplorhini: Tarsioidea. New
York: Interscience.

Lee-Thorp, Julia, J. Francis Thackerary, and Nikolaas van der Merwe. 2000. “The Hunters and the Hunted
Revisited.” Journal of Human Evolution 39 (6): 565–576. https://doi.org/10.1006/jhev.2000.0436.

McKee, Jeffrey K., Frank E. Poirier, and W. Scott McGraw. 2005. Understanding Human Evolution. 5th ed.
Upper Saddle River, NJ: Pearson Prentice Hall.

Meier, Joana I., David A. Marques, Salome Mwaiko, Catherine E. Wagner, Laurent Excoffier, and Ole Seehausen.
2017. “Ancient Hybridization Fuels Rapid Cichlid Fish Adaptive Radiations.” Nature Communications 8.
https://doi.org/10.1038/ncomms14363.

Phillips, Melissa Lee. 2006. “Genetic evidence for punctuated equilibrium.” The Scientist, October 5, 2006.
https://www.the-scientist.com/daily-news/genetic-evidence-for-punctuated-equilibrium-47144.

Pruetz, J.D., et. al, P. (2015). New evidence on the tool-assisted hunting exhibited by chimpanzees (Pan
troglodytes verus) in a savannah habitat at Fongoli. R. Soc. open sci.2140507140507http://doi.org/10.1098/
rsos.140507

Stringer, Chris, and Peter Andrews. 2011. The Complete World of Human Evolution. 2nd ed. London: Thames &
Hudson.

Canine sexual dimorphism in Ardipithecus ramidus was nearly human-like. Proceedings of the National
Academy of Sciences, 118 (49) e2116630118; DOI: 10.1073/pnas.2116630118

Takahashi, Tetsumi, and Stephan Koblmüller. 2011. “The Adaptive Radiation of Cichlid Fish in Lake
Tanganyika: A Morphological Perspective.” International Journal of Evolutionary Biology 2011.
https://doi.org/10.4061/2011/620754.

Wayman, Erin. 2011. “How Africa Became the Cradle of Humankind.” Smithsonian, October 17, 2011.
https://www.smithsonianmag.com/science-nature/how-africa-became-the-cradle-of-
humankind-108875040/.

Access for free at openstax.org


CHAPTER 5
The Genus Homo and the Emergence
of Us

Figure 5.1 Liang Bua Cave on the island of Flores in Indonesia. A potentially new species of the genus Homo,
Homo floresiensis was discovered in this cave in 2003. (credit: “Flores: Ruteng to Bajawa” by Bryn Pinzgauer/flickr,
CC BY 2.0) (credit: “Flores” by Ryan Somma/flickr, CC BY 2.0)

CHAPTER OUTLINE
5.1 Defining the Genus Homo
5.2 Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus
5.3 The Emergence of Us: The Archaic Homo
5.4 Tracking Genomes: Our Human Story Unfolds

INTRODUCTION Our human story continues with the rise of the genus Homo which at one time represented
at least 8 different species in our human lineage – with only H. sapiens surviving. The genus Homo displays
some of the most diverse and complex examples of both australopithecine and Homo characteristics, which
has made the classification of species in this genus challenging. In this chapter we take a look at how
paleoanthropologists have defined Homo and at attempts to answer the question, “What does a species of the
genus Homo look like?”
142 5 • The Genus Homo and the Emergence of Us

5.1 Defining the Genus Homo


LEARNING OBJECTIVES

By the end of this section, you will be able to:


• Describe the time periods and geological context of the genus Homo.
• Identify some key differences between the genus Homo and Australopithecus.
• Define some of the limitations of and challenges in the classification of hominin species in the genus Homo.
• Explain the concept of encephalization as it relates to early hominin evolutionary development and as a tool for
hominin classification.

Putting Homo into Context


Before learning about the hominin species that make up the category genus Homo, it will be helpful to become
familiar with the key archaeological time periods with which Homo is associated. The species and cultural
developments mentioned below will be explored in greater detail in the sections that follow.

• Lower Paleolithic (from roughly 3 million years ago to approximately 300,000 BCE): This period includes
H. habilis and the Oldowan tool industry, followed by H. ergaster and the Acheulean tool industry.
• Middle Paleolithic (approximately 300,000–40,000 BCE): This period includes continued use of
Acheulean tools by H. heidelbergensis, followed by H. neanderthalensis and the Mousterian tool industry.
• Upper Paleolithic (approximately 43,000–26,000 BCE): The Upper Paleolithic saw the emergence of cave
art like that found in the famous Chauvet Cave in France (Figure 5.29), Venus figurines (Figure 5.28), and
an increased use of bone and antler in tools and jewelry. The most recent ice age occurred during this
time, with glaciers covering huge parts of the planet. The emergence of Paleoindians and the use of Clovis
points, which were used to kill large game such as mastodons and mammoths, occurred near the end of
this time period.
• Neolithic (Agricultural Age) (8,000–3,000 BCE): New innovations appear during the agricultural age, or
“Neolithic revolution,” as H. sapiens set up permanent settlements. Humans begin to transition from
being hunters and gatherers to growing grow crops, owning land, and domesticating animals.

The Challenge of Defining the Genus Homo


The previous chapter introduced the australopithecines, who were diverse in their physical characteristics
(gracile and robust), with large jaws and teeth and small brain size. A key characteristic shared by both the
australopithecines and the genus Homo is bipedalism. The transition to bipedalism is linked with various
anatomical changes, including longer legs, changes in spinal curvature, and the development of arches in the
feet to conserve energy and increase balance when walking.

What criteria other than bipedalism might be used to classify a species under the genus Homo? Many
anthropologists have attempted to establish specific criteria to use in determining a classification of Homo.
Paleoanthropologists Mary Leakey, Louis Leakey, and John Napier, as well as primatologist Phillip Tobias,
were among the first to extensively study the fossils of Homo habilis, considered to be one of the earliest
species in the genus Homo. H. habilis had a brain size of around 661–700 cc, which was larger than the
australopithecines’, with hands that were capable of the dexterity needed for making tools, due to bone
structure changes and a repositioning of the thumb, which allowed for better grip.

The type specimen OH 7 of H. habilis dated between 2 and 1.7 MYA and was found in 1960 at Olduvai Gorge by
Jonathan and Mary Leakey. It was described by Louis Leakey in 1964. Type specimen refers to a specimen that
serves as the standard for the taxon or classification group for that species. OH 7 is the identification or
accession number of this specific specimen and stands for “Olduvai Hominid #7.” The specimen consisted of a
partial juvenile skull, hand, and foot bones. It possessed teeth that were much smaller than those of any
australopithecine and was possibly in coexistence with the robust australopithecines (Paranthropus). Based on
an endocranial cast (an imprint of the interior of the brain case), it was determined that H. habilis may have
possessed what is called a Broca’s area in the brain. Broca’s area, which includes two Brodmann areas
(referred to as 44 and 45), is located in the middle of the left cerebral cortex of the brain and is especially
important for speech development (Figure 5.2). Some scientists have suggested that H. habilis started to

Access for free at openstax.org


5.1 • Defining the Genus Homo 143

develop the neural networks necessary for human speech, while others argue that H. habilis probably already
had speech.

FIGURE 5.2 Position of the Broca’s area in the brain, consisting of Brodmann areas 44 (yellow) and 45 (blue).
Broca’s area is associated with speech development and may have been present in the brain of H. habilis. (credit:
Fatemeh Geranmayeh, Sonia L. E. Brownsett, Richard J. S. Wise/Wikimedia Commons, CC BY 3.0)

The postcranial features (skeletal structures in the body other than the skull) of Homo habilis are not as well
established, as is the case for many other early hominin fossils. This can be problematic, as many hominin
species coexisted with overlapping traits. Likewise, it can be problematic to have postcranial material and not
the cranium or skull. The skull often serves as a diagnostic tool when postcranial materials do not provide
enough evidence or provide confusing evidence.

Based on their research on H. habilis, Mary Leakey, Louis Leakey, and John Napier proposed the following
criteria for classifying Homo: a brain size over 600 cc; a round, globular skull; tool use; reduced prognathism
(protrusion of the jaw) and smaller jaws and mandibles; humanlike postcranial features; and feet that are fully
adapted for walking (Leakey, Tobias, and Napier 1964). While this list established specific and fairly
comprehensive guidance, the diversity of traits and the ways in which they overlapped didn’t always line up
with the criteria.

H. habilis has been at the center of several debates regarding their taxonomic position and relationship with
other early archaic Homo species. For example, H. habilis was initially believed to have been a direct human
ancestor through the lineage of Homo erectus and then modern humans. This viewpoint is now debated and
has resulted in a scientific divide between those supporting H. habilis and those suggesting another Homo
species, H. rudolfensis, as being the ancestor of H. erectus. H. rudolfensis is an archaic Homo dated to about 2
MYA, which coexisted with other Homo species during that time period. A cranium was discovered in 1972
along Lake Turkana in Kenya by Bernard Ngeneo, a local Kenyan. The specimen was later described by
paleoanthropologist Richard Leakey. There is a lot that is not known about this species; scientists are missing
postcranial materials, and as of yet no tools have been found. There are hypotheses that propose that H.
rudolfensis might be a H. habilis male, exhibiting a larger cranium than that seen in a female H. habilis. Others
suggest it is a completely different species. Another controversy centers on tool use. While Homo habilis was
long regarded as the earliest hominin to use stone tools, it has been determined, based on evidence of
cutmarks, that at least one australopithecine (A. garhi) used stone tools before H. habilis, at around 2.6 MYA
(Semaw et al. 1997).
144 5 • The Genus Homo and the Emergence of Us

FIGURE 5.3 The specimen of H. rudolfensis on the left is noticeably different from that of H. habilis on the right.
(credit: Conty/Wikimedia Commons, CC BY 3.0)

While there are still questions as to the phylogenetic relationship of H. habilis and H. rudolfensis, there is
general agreement that Homo did evolve from Australopithecus. The timing and placement of the split
between Australopithecus and Homo, however, is still debated. H. habilis was determined to not be an
Australopithecus due to its smaller teeth, a humanlike foot, and hand bones that suggested an ability to
manipulate objects with precision.

One of the main considerations in classifying H. habilis as a Homo and not an Australopithecus was its cranial
capacity, which is a measurement that indicates brain size. With some exceptions, cranial capacity can serve
as an indicator of where a hominin fossil might belong in the hominin phylogenetic tree. Encephalization
refers to a progressive increase in brain size over time. In human evolution, we can observe encephalization
beginning with Homo habilis and progressing more rapidly through H. erectus. Encephalization correlates
with an increase in behavioral, cognitive, and cultural complexity. Cognitive developments correspond with
our ability to construct and form ideas, including the ability think in and communicate via symbolic and
abstract language, such as that used in storytelling, ritual, and art. There are always exceptions, however, such
as the island-dwelling, small-brained H. floresiensis, who will be introduced later in this chapter. In spite of
having a very small brain, H. floresiensis made and used tools and built fires. This discovery has challenged
what we thought we knew about the correlation of brain size and cognitive development in human evolution.

The encephalization quotient (EQ) can serve as a good indicator (with some exceptions) of classification
within the genus Homo. The encephalization quotient is a calculation arrived at by comparing the ratio
between actual brain size (determined with either a mass or volume calculation) and expected brain size. Body
size is a factor in these measurements as expected brain size reflects the relationship between brain and body
size for a given taxonomic group (Jerison 1973). The larger the brain weight relative to the overall body weight,
the more likely that the brain was used for more complex cognitive tasks. Harry J. Jerison (1973) was the first
to develop EQ measurements. The formula he used for calculating EQ in birds and mammals is brain mass/
0.12 × (body mass)0.66. Other formulas have also been proposed, such as EQ = brain mass (11.22 × body mass
0.76) (Martin 1981). While EQ is a strong tool for studying brain size in early hominins, there are always
potential margins of error when dealing with fragmentary fossils, and increasingly alternative forms of
measurements are being proposed. One study proposes that EQ should no longer be used as a tool in
calculating brain size in primates and other vertebrate species, based on the premise that cognitive
performance does not depend on body size and so body size should not be included in the formula (Schaik et
al. 2021). Other theories consider the number of cortical neurons and neural connections as most important
when considering cognitive ability (Roth and Dicke 2012). According to this approach, the density of the cortex

Access for free at openstax.org


5.2 • Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus 145

is more associated with intelligence than is brain size. These alternate approaches would perhaps better
explain those exceptions in the fossil records, such as H. floresiensis. Other interesting research is looking at
potential levels of cognition and memory as it relates to levels of tool complexity (Read and van der Leeuw
2008).

In spite of these criticisms, many see EQ measurements as providing fairly consistent results. Modern humans
(Homo sapiens) have an EQ of roughly 6.0–7.0 (meaning that their brain mass is six to seven times greater
than what one would expect to find in a comparable mammal of the same body size). H. erectus has an EQ of
4.0, while for an australopithecine EQ is around 2.5 to 3.0 (Fuente 2012, 227). Figure 5.4 shows increases in
average brain sizes for various species over time.

FIGURE 5.4 After remaining steady for millennia, average brain size increases noticeably in the last two million
years. (Xiujie, Wu, and Norton 2007). (credit: Gisselle Garcia, artist (brain images), CC BY 4.0)

5.2 Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus
LEARNING OBJECTIVES

By the end of this section, you will be able to:


• Compare and contrast the anatomy and material culture of H. habilis, H. erectus, and H. ergaster.
• Define the term “tool industry” and describe the tools typified by the Oldowan and Acheulean industries.
• Identify possible correlations between the environment, diet, new behaviors, and brain growth.

The Toolmakers
Archeologists use the word industry to describe a classification or assemblage of stone tools. The Oldowan tool
industry is the oldest known stone tool industry. It dates from around 2.5 to 1.5 MYA. Because there were
several hominins in Africa during this time, it is unclear whether these tools were created and used by H.
habilis or by Paranthropus boisei, or by both (Susman 1991). Oldowan tools are fairly crude and primitive in
appearance, which can make it difficult to find and identify them in the field.
146 5 • The Genus Homo and the Emergence of Us

FIGURE 5.5 An Oldowan tool. This chopper is made of quartzite and dated to the lower Paleolithic period. (credit:
Locutus Borg/Wikimedia Commons, Public Domain)

Mary Leakey was the first to create a system to classify Oldowan assemblages, basing her classification on
utility, or how the tools were used. Later efforts were made to classify tools based on how the tools were made.
All Oldowan tools were created by using hard hammer percussion, in which flakes are chipped away from a
stone, resulting in a “core”. These cores served as basic tool that could have been used for killing game, cutting
meat and plants, and possibly woodworking. Oldowan toolmaking is the earliest evidence of “flint knapping,” a
technique that became more complex over time, resulting in more sophisticated tools (Figure 5.6).

FIGURE 5.6 Demonstration of flint knapping, an ancient technique for shaping stones into useful tools. (credit:
“Flint-knapping Demonstration” Tonto National Monument/NPS photo/flickr, CC BY 2.0)

Handedness, or brain lateralization (i.e., whether one is right-handed or left-handed), is a cognitive


development that can be inferred through evidence of the use of a dominant hand in creating and using tools.
The use of a dominant hand suggests a possible reorganization of the brain. It is believed that about 90 percent
of humans are right-handed, which differs from apes, which are closer to 50 percent. David Frayer (2016), an
anthropologist from the University of Kansas, has concluded that the brain lateralization of Homo habilis was
more like that of modern human than that of apes. Frayer found striations on the teeth of a 1.8-million-year-
old Homo habilis fossil that indicate right-handedness. He concluded that meat was pinched between the teeth
and held in place with the left hand, while the right hand cut the meat with a tool. Brain lateralization,
increasing brain size, and tool use are just some of the key developments we see in the genus Homo.

Access for free at openstax.org


5.2 • Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus 147

Homo ergaster
Homo ergaster is the first Homo that looks much like H. sapiens. A key difference between H. ergaster and
earlier hominins is that H. ergaster exhibits substantially less sexual dimorphism in body size. H. ergaster
males were only 20 percent larger than females. Likewise, modern human males are only 15 percent larger
than females. This contrasts sharply with all other previous hominins, such as the australopithecines, in which
males were 50 percent larger than females. It is well established that in mammals, significant dimorphism is
associated with polygyny, and a lack of dimorphism is associated with a monogamous mating system. It has
been suggested that the reduction in dimorphism seen in H. ergaster may indicate less male-male competition
for access to females and perhaps a shift toward a monogamous mating system, with substantial parental
investment in offspring.

Other similarities between H. ergaster and modern humans are seen in the teeth and postcranial features. The
average cranial capacity of H. ergaster is 1,100 cc, which is just a bit smaller than that of modern humans, who
average 1,400 cc. There is a very important specimen of H. ergaster that bears mentioning, the Nariokotome
Boy. This specimen was discovered in 1984 by Kenyan paleontologist Kamoya Kimeu near Lake Turkana in
Kenya. It is dated to approximately 1.6 MYA. The specimen is believed to represent a boy of about 12 years old,
determined by various dental and cranial features. He was about 5 feet 4 inches tall, roughly the same height
as a modern boy of the same age (Figure 5.7). It has been estimated that his adult height would be around 5 feet
10 inches, with an estimated cranial capacity of 900 cc. The Nariokotome Boy looks tremendously modern in
appearance despite being 1.6 million years old.

FIGURE 5.7 This specimen of Homo ergaster is known as the Nariokotome Boy. It is believed to be the remains of
a boy who was about 12 years old at the time of death. (credit: “Homo ergaster (fossil hominid) (Lower Pleistocene,
1.5 to 1.6 Ma; Nariokotome, Lake Turkana area, Kenya) 4” by James St. John/flickr, CC BY 2.0)

Homo ergaster Technology


Homo ergaster continued to use Oldowan stone tools, but they also began to construct much more complex
tools, referred to as the Acheulean industry (Figure 5.8). These tools have been found throughout Africa,
Europe, and the Middle East and are first noted as appearing approximately 1.6 MYA to 200,000 years ago.
These types of tools are rarely found in Asia. It is currently unclear whether this is because the Acheulean
industry had not yet been developed when H. erectus migrated to Asia or because bamboo (a plant found in
148 5 • The Genus Homo and the Emergence of Us

abundance in Asia) was found to be a more versatile resource than stone. As wood and bamboo are
biodegradable, no remains of tools constructed from these materials would exist today.

FIGURE 5.8 This hand axe, found in the Zamora province, Spain, displays the form and construction techniques
typical of the Acheulean industry. (credit: Jose-Manuel Benito/Locutus Borg/Wikimedia Commons, Public Domain)

Unlike Oldowan tools, Acheulean tools actually look like tools. Acheulean tools are distinct from Oldowan tools
in that they were modified on both sides, resulting in a symmetrical tool with two faces, also known as biface.
One end of the tool was tapered, while the other end was rounded. The creation of symmetrical objects from
stone materials is believed to represent an increase in cognitive ability as well as motor skills in the tool maker.
These bifaces were struck from large flakes, which had themselves been struck from boulder cores. This
required a more delicate technique than banging one rock into another. Acheulean tools were typically created
used the soft hammer technique. In this technique, hard rock such as flint is chipped by striking it with a softer
material such as bone or wood. The gentler blows detach small flakes that leave smooth, shallow scars,
creating a straighter and more uniform cutting edge.

The main advantage of Acheulean technology is that it allowed hominins to get a better grip on their tools, as
they were shaped to fit the hand. This tool type was used primarily as a hunter’s knife but also for chopping,
scraping, and even piercing. The most common type of biface tool is a hand axe. Note that even though these
tools are called axes, they are held in the palm of the hand. Another type of Acheulean biface used by Homo
ergaster is called a cleaver (Figure 5.9). The cleaver had a wide cutting edge across the end instead of a point
and was best suited for hunting or hacking wood. Another Acheulean tool is the side scraper, used to scrape
hides that could then be turned into simple clothing.

Access for free at openstax.org


5.2 • Tools and Brains: Homo habilis, Homo ergaster, and Homo erectus 149

FIGURE 5.9 Map of Acheulean cleaver finds dated to the Lower Paleolithic (1.76–0.13 MYA). Note the
concentration of artifacts found in certain areas of Africa and in Spain. (attribution: Copyright Rice University,
OpenStax,under CC BY 4.0 license)

Evidence of an Increase in Meat Eating


In 1973, a specimen of H. ergaster known as KNM ER 1808 was found in Koobi Fora, Kenya. Dated to about 1.7
MYA, this is the most complete H. ergaster specimen ever found. Analysis of KNM ER 1808 suggests that H.
ergaster may have been eating carnivore liver, which is high in Vitamin A. This may indicate a dietary shift
toward increased meat eating by H. ergaster.

Homo erectus: A Success Story


Homo erectus is the longest-surviving species in the genus Homo. For almost two million years, H. erectus
existed and evolved. Also known as the “Upright Man” or Java Man, H. erectus was first found in Indonesia in
1891 by Eugene Dubois, a professor of anatomy at the University of Amsterdam. At a site called Trinil, he found
a skull cap and a femur. He named the specimen Pithecanthropus erectus. The most current dates for Homo
erectus are 1.2–1.6 million years ago. H. erectus exhibits a cranial capacity averaging 900 cc and several
distinguishing characteristics. These characteristics include a slightly projecting nasal spine, shovel-shaped
incisors, a nuchal crest (a ridge in the back of the skull that supported strong neck muscles), very thick skull
bones, and pronounced brow ridges. They also had longer legs, evidence that they were utilizing energy much
more efficiently when walking and becoming effective hunters. We also see a diminishing of the protruding jaw
(or prognathism) that was so prominent in the australopithecines.
150 5 • The Genus Homo and the Emergence of Us

FIGURE 5.10 This Homo erectus cranium exhibits a number of defining features, including a projecting nasal
spine, thick skull bones, and pronounced brow ridges. (credit: Daderot/Wikimedia Commons, Public Domain)

There is evidence that H. erectus was using fire around 1.7–2.0 MYA, which would make it the first or one of
the first hominins to do so. Ancient hearths, charcoal, and charred animal bones have been found in
Zhoukoudian, China. This evidence suggests that H. erectus was hunting, cooking, and eating meat. Also found
at Zhoukoudian are a number of fossil skulls that were once thought to display evidence of cannibalism.
However, recent research evidence suggests that the remains of these H. erectus were prey to animal
scavengers such as hyenas (Boaz et al. 2004).

The Smithsonian Institution has created an interactive tool (https://openstax.org/r/human-evolution-


interactive-timeline) that visually illustrates the interrelationships between an increasingly variable and
colder climate, encephalization, bipedalism, and new technologies and tool use. These correlations align with
fossil evidence indicating changes in diet and caloric requirements in response to a colder and changing
climate, which ultimately fueled a growing brain. The “expensive tissue hypothesis” proposes that maintaining
a brain is metabolically expensive and that, in order to meet the energy requirements of a larger brain, our
digestive system became smaller and shorter, making it more suited for higher-quality, nutrient-dense food
such as meat (Aiello and Wheeler 1995). The list below summarizes some of the key evolutionary changes seen
in H. erectus from 2 MYA to possibly as recent as 50,000 years ago, which provide further support for these
correlations (Dorey 2020).

1. There is a progressive increase in brain size in H. erectus, from about 550 cc to 1,250 cc.
2. There is evidence of increased use of fire and of eating cooked meat at H. erectus sites. H. erectus would
have needed as much as 35 percent more calories than previous hominins (Fuentes 2012).
3. The eating of softer foods as a result of cooking meat and plants alleviated the need for large chewing teeth
and jaws. Over time teeth became smaller, which resulted in thicker enamel.
4. There is a gradual decrease in prognathism, and as in H. habilis, skulls provide evidence of smaller teeth
and jaws, which would have made room for larger brains.
5. H. erectus is taller than any other previous hominin, with longer legs that provided the ability to run great
distances and chase prey. New research is shedding some additional light on the possible benefits of
running in early hominins The fossil evidence suggests that endurance running is a derived adaptation of

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 151

the genus Homo, originating about two million years ago, and may have been instrumental in our
evolution (Bramble and Lieberman 2004).

The Homo ergaster and Homo erectus Debate


There is great debate as to whether Homo ergaster and Homo erectus are one species or two. Some refer to H.
ergaster as the “early” H. erectus. Their differences are largely geographical: H. ergaster is associated with
Africa and H. erectus with Asia. Yet some researchers have concluded that H. ergaster and even H. habilis
should be referred to as H. erectus. Whether to lump or split the diverse species in the genus Homo is an
ongoing challenge in the scientific community. While there are some anatomical differences between H.
erectus and H. ergaster, they are fairly minimal.

FIGURE 5.11 Homo erectus (left) has a sagittal keel (ridge on top of head), a shorter forehead, and a different-
shaped skull than Homo ergaster, seen on the right. (credit: (left) kevinzim/Wikimedia Commons, CC BY 2.0; (right)
Reptonix free Creative Commons licensed photos/Wikimedia Commons, CC BY 3.0)

The diversity and number of evolutionary changes seen in H. erectus indicate that H. erectus set the stage for
the arrival of the archaic Homo, which we will cover in the next section.

5.3 The Emergence of Us: The Archaic Homo


LEARNING OBJECTIVES

By the end of this section, you will be able to:


• Describe the context, time frame, and key anatomical characteristics of archaic Homo.
• Explain the potential environmental conditions that led to evolutionary change in anatomy and material culture
seen in archaic Homo.
• Compare and contrast the current hypotheses regarding the extinction of the Neanderthal.

Defining the Archaic Homo


There is no universal consensus on what is included within the term “archaic Homo.” The term is used as an
umbrella category encompassing all the diverse Homo species after H. erectus. Hominin species classified as
archaic Homo typically have a brain size averaging 1,200 to 1,400 cc, which overlaps with the range of modern
humans. Archaic Homo are distinguished from anatomically modern humans by the characteristics of a thick
skull, prominent supraorbital ridges (brow ridges), and lack of a prominent chin. Archaic Homo are viewed as
transitional between H. erectus and H. sapiens and display many overlapping and varied traits. It has been
proposed that archaic Homo may have been the first species to use language, based on the size of their brains
152 5 • The Genus Homo and the Emergence of Us

and the fairly large social groups they lived in. Archaic Homo species as presented here will be divided into
two groups: the Early Archaic (800–250 KYA) and the Late Archaic (300–30 KYA).

Early Archaic Homo


Homo antecessor
Homo antecessor has been found in Spain, France, and England and dates to around 1.2 MYA to 800 KYA.
These specimens represent the oldest fossil evidence for the presence of the genus Homo in Europe. Some
scientists have suggested that this species is the ancestor of Homo heidelbergensis, while others suggest that
H. antecessor is the descendent of H. ergaster. Homo antecessor was first found at the Sima de los Huesos site
of the Sierra de Atapuerca region in Spain. Within this site is a cave known as the Pit of Bones, where more
than 1,600 fossils of 28 individuals have been found that date at or before 780,000 years ago. The site is an
important one that stretches over a long period of time and displays the emergence and divergence of various
Homo physical characteristics that later appear in the Neanderthal. Evidence from nuclear DNA suggests that
early hominins at this site were related to the Neanderthal and not the Denisovans, indicating divergence
earlier than 430,000 years ago (Meyer et al. 2016). The section on the Neanderthal will explore further the
interbreeding and divergences of the Neanderthal, Denisovans, and modern Homo sapiens.

Homo antecessor was almost six feet tall and males weighed about 200 pounds, well within the range of
variation for modern humans. Other anatomical features of this species include a protruding occipital bun (a
bulge found in the occipital area of the skull), a low forehead, no strong chin, and a cranial capacity of about
1,000 cc. It has been suggested that the purpose of the occipital bun is to balance the weight of the anterior
portion of the skull and face. One very modern trait exhibited by this species is the presence of a facial
depression above the canine tooth called the canine fossa, which is also found in modern humans. The best-
preserved fossil is a maxilla (upper jawbone) of a 10-year-old individual.

In addition to the fossil bones, 200 stone tools and 300 animal bones were also found at Gran Dolina, another
location at the Atapuerca site, along with a carved stone knife. Stone tools at this site were predominantly
Oldowan style and constructed from local raw materials. Tools included cutting flakes and hand-held cores. It
has been suggested that the absence of retouched tools at this site indicates that these tools were created
primarily for processing and eating meat. Cutmarks are present on the majority of animal remains. One of the
most intriguing observations about this site is that there are numerous large animal carcasses (mostly deer)
that are believed to have been transported to the site rather than consumed where they were killed. Some
scientists have suggested that the practice of bringing food back to the site is evidence of social cooperation,
suggesting both a division of labor and a custom of food sharing.

Many of the bones of Homo antecessor show the same evidence of cutmarks as the animal bones, indicating
that flesh was removed from the bones with the goal of dismemberment. Some scientists have taken this to
mean that H. antecessor practiced cannibalism. However, humans have also been known to remove the flesh
from bones during funerary rites. Whether the cutmarks made by H. antecessor represent cannibalism, a
funerary rite, or another yet unknown practice is still being debated.

Homo heidelbergensis
Homo heidelbergensis is an incredibly variable group. Many archaic Homo species are included in this group
because they possess features that can best be described as a mosaic between H. ergaster, H. erectus, and
anatomically modern humans (AMH). This section looks at just a few of the specimens that are regularly
attributed to Homo heidelbergensis.

One of the most important Homo heidelbergensis specimens is known as Mauer. It was found in 1907 in
Germany and is represented by a mandible (lower jaw) that is dated to approximately 600,000 years ago. It has
a robust mandible and a receding chin like earlier Homo ergaster but has very small molars like anatomically
modern H. sapiens. The jaw is so big and the teeth are so small that there is plenty of space for additional teeth
to develop behind the wisdom teeth. Given that the third molar (the wisdom tooth) has already erupted, it has
been suggested that this individual was between 20 and 30 years at death.

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 153

FIGURE 5.12 This jawbone from a Homo heidelbergensis specimen was found in Germany in 1907 and is dated to
approximately 600,000 years ago. (credit: Gerbil/Wikimedia Commons, CC BY 3.0)

Another important specimen of Homo heidelbergensis is known as the Petralona cranium. It was found in
1960 in Greece. Dates are uncertain but believed to be in the range of 100,000 to 700,000 years. Animal fossils
found with the specimen indicate Petralona is between 350,000 and 200,000 years old. It combines Homo
ergaster–like traits, such as massive brow ridges and thick cranial bones, with a cranial capacity of 1,200 cc,
which is similar to anatomically modern H. sapiens.

A third specimen of Homo heidelbergensis is known as Bodo. It is very possibly the oldest archaic human
specimen from Africa and was found in Ethiopia in 1976. It is dated to approximately 600,000 years and has a
relatively large cranial capacity of 1,250 cc, which is again within the range of variation for modern humans. It
is a robust cranium with very thick bones and two separate brow ridges.

Homo heidelbergensis Technology and Culture


Bodo is associated with Acheulean bifacial hand axes. Some scientists have suggested that Bodo butchered
animals because Acheulean hand axes have been found with animal bones. There are cutmarks on the Bodo
cranium that resemble those made by cutting fresh bone with stone tools. It has been suggested that the Bodo
cranium is the earliest evidence of the removal of flesh immediately after death using a stone tool. The
cutmarks were made symmetrically and with specific patterns on the cranium, which is interpreted as strong
evidence that the defleshing was done purposefully for funerary practices. Once again, others have suggested
that the cutmarks indicate that Bodo may have been practicing cannibalism.

In addition to their use of stone tools from the Acheulean tool industry, Homo heidelbergensis is also believed
to have used spears. The earliest known spears have been found in Schöningen, Germany, and are dated to
about 400,000 years ago. The spears were made either from spruce or pine wood and are believed to have had
a range of about 35 meters. Probably the most important technological achievement evident in these spears is
the use of hafting technology. Hafting involves attaching stone points to a handle made of another substance,
such as wood, metal, or bone. The spears found at Schöningen represent one of the first known instances in
which hominins united separate elements into a single tool.

Hafting gives stone tools more utility, as they can now be thrown (as with a spear), shot (as with an arrow), or
used with more leverage (like an axe). These hafted stone points are able to be used with increased force and
effectiveness, allowing people to hunt and kill animals more efficiently. This increased efficiency in hunting
and killing animals is believed to have created a situation in which H. heidelbergensis had regular access to
154 5 • The Genus Homo and the Emergence of Us

meat and other high-quality foods. Some have suggested that the presence of spears represent evidence that
H. heidelbergensis could hunt herd animals that can run faster than a human, and that they had sophisticated
hunting strategies requiring cognitive skills like anticipatory planning.

Like Homo ergaster and Homo erectus, Homo heidelbergensis occupied both caves and open-air sites.
However, they did not just use the sites as is, they modified them. One of the most interesting aspects of the
cultural behavior of Homo heidelbergensis is that they are associated with clear archeological evidence for
modified dwellings. For example, in the Czech Republic there is a modified dwelling that consists of a stone
foundation that is approximately 700,000 years old. Most likely, this dwelling had a roof constructed of thick
branches. Other modified dwellings have been found in Germany and France.

Evidence of controlled fire has been found at most reasonably preserved Homo heidelbergensis sites. The
oldest established continuous fire site for Homo heidelbergensis is from Israel and is dated to around 780,000
years old.

FIGURE 5.13 Phylogenetic tree and proposed migration routes of genus Homo heidelbergensis and later
Denisovans and Neanderthals. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

Late Archaic Homo


Homo naledi: A Rising Star
The most recently described archaic Homo is known as Homo naledi. They were found in the Rising Star cave
system in South Africa in 2013 and 2014 (Figures 5.14–5.15) and are dated to approximately 235,000–335,000
years old. Over 1,500 bones from as many as 15 individuals were recovered from the cave, which is possibly
the largest assemblage of a single hominin species yet discovered. Despite their relatively recent date, they
have exceptionally small cranial capacities, comparable to the robust and gracile australopithecines, which are
around 560 cc. The encephalization quotient of H. naledi is estimated at 4.5, which is the same as H.
floresiensis but notably smaller than all other Homo (contemporary Homo are all above 6). The presence of
this small-brained hominin at the same time that Neanderthals and Homo heidelbergensis were around is
further evidence that multiple hominin lineages were coexisting and evolving at the same time. The
classification of H. naledi proved to be a challenge, as the specimens presented a mosaic of traits and
characteristics associated with an array of other hominin species.

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 155

FIGURE 5.14 Maps showing the location of the Cradle of Humankind World Heritage Site in South Africa, where
Homo naledi fossils were found in the Rising Star cave system. (credit: Hawks et al. (2017), eLife, CC BY 4.0)

FIGURE 5.15 The Rising Star cave system, showing geological features and the location of the excavation area
where numerous Homo naledi specimens have been found. (credit: Paul H. G. M. Dirks et al. (2015),
eLife/Wikimedia Commons, CC BY 4.0)
156 5 • The Genus Homo and the Emergence of Us

FIGURE 5.16 H. naledi skulls. It is apparent in these images that this species had rather pronounced prognathism
(Credit: John Hawks, Marina Elliott, Peter Schmid et al. (2017), eLife/Wikimedia Commons, CC BY 4.0)

FIGURE 5.17 H. naledi feet were much like those of modern humans. (credit: W. E. H. Harcourt-Smith, Z.
Throckmorton, K. A. Congdon, B. Zipfel, A. S. Deane, M. S. M. Drapeau, S. E. Churchill, L. R. Berger & J. M. DeSilva
(2015)/Nature Communications/Wikimedia Commons, CC BY 4.0)

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 157

FIGURE 5.18 The hands of H. naledi display curved finger bones and large thumbs, indicating that it still had an
adaptation for climbing trees. (credit: Lee R. Berger et al. (2015), eLife/Wikimedia Commons, CC BY 4.0)

FIGURE 5.19 Comparison of some of the most commonly known Homo species (credit: Chris Stringer, Natural
History Museum, United Kingdom (2015), eLife/Wikimedia Commons, CC BY 4.0)

Homo naledi: Did They Bury Their Dead?


Homo naledi has not yet been found in association with any stone tools. Despite a lack of established tool use,
there is fairly convincing evidence that H. naledi may have used the cave system as a place to bury their dead.
The hypothesis that H. naledi had a ritualistic mortuary practice is based on several observations, such as the
bones appearing to lack evidence of gnawing marks from predators and the lack of evidence of layers of
sediment that would suggest the bones were deposited by flooding (Dirks et al. 2015). In 2017 additional fossil
remains were found in a second chamber in the Rising Star cave system (Hawks et al. 2017), but these remains
don’t as yet appear to offer additional evidence to support the hypothesis of an intentional burial.

Some scientists believe that there is insufficient evidence to conclude that H. naledi were involved in funerary
ritual practices. They have noted that the preservation of H. naledi specimens are similar to that of cave-
dwelling baboons that have died natural deaths. At Sima de los Huesos, remains of about 28 Neanderthal and
H. heidelbergensis fossils were found in a cave dated to about 430,000 years ago. Researchers who examined
the scattering patterns of the remains at both the Rising Star cave system in Africa and the Sima de los Huesos
site in Spain (Egeland et al. 2018) concluded that the sites showed evidence of having been scavenged but that
this doesn’t disprove the possibility that they may also be deliberate burials. The verdict is still out on this. Lee
158 5 • The Genus Homo and the Emergence of Us

Berger and other scientists are conducting further investigations of the H. naledi skeletal deposits to further
explore the possibility they might be evidence of something more deliberate than the actions of predators.

Rethinking the Neanderthal


Homo neanderthalensis
The word “Neanderthal” might conjure up stereotypical images of a brutish caveman-like creature holding a
club in one hand and dragging supper with the other. No one said entertainment had to be scientifically
accurate, but media can create false perceptions and stereotypes about the past. This section takes a closer
look at who the Neanderthal people were and the role they played in the human story.

FIGURE 5.20 Distribution map of Neanderthal sites. The red squares mark locations of Neanderthal remains and
the shaded area represents the supposed territory of Neanderthal people in Europe and Asia. (credit: modification
of work Neanderthal distribution by Berria/Wikimedia Commons, CC BY-SA 4.0)

Neanderthals have been found only in regions of Europe and the Middle East and are dated to between about
400,000 and 40,000 years ago. The first fossils, which were found in the Neander Valley, were believed to be
the remains of an extinct kind of human. The Germans called them the Neanderthals, the people of the
Neander Valley.

FIGURE 5.21 The Neanderthal skull on the left is noticeably different from the H. sapiens skull on the right. (credit:
(left) Jose-Manuel Benito/Locutus Borg/Wikimedia Commons, Public Domain; (right) “Image from page 27 of
“Human physiology” (1907)” by Furneaux, William S/Internet Archive Book Images/flickr, Public Domain)

Neanderthals possess several distinctive anatomical characteristics: the skull and brain is larger than that of

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 159

humans, with an average size in Neanderthals of 1,520 cc compared to modern humans’ 1200–1400 cc. Does
the Neanderthal’s larger brain size mean that it was more intelligent than modern humans? As mentioned
earlier in this chapter, while there does seem to be a correlation between brain size and complex cognitive
skills, the brain in some hominins may have been organized differently than that of modern humans, with
different anatomical areas of the brain emphasized. It is believed that in the Neanderthal brain, the frontal
region, which is the center of speech and language, was less developed, while the back of the brain, which
deals with the senses, was more developed. This greater development in the back area of the brain could be a
survival adaptation found in Neanderthals who had to hunt in often harsh and difficult conditions.

Philip Lieberman, a cognitive scientist at Brown University, argues that Neanderthals lacked the anatomy
necessary for humanlike speech. He drew this conclusion based on a reconstruction of a Neanderthal throat,
which indicated that the neck could not accommodate the vocal apparatus of modern humans (Lieberman,P.
2007). While there is evidence of a hyoid bone, a small horseshoe-shaped bone in the front of the neck, that
would have been able to anchor the tongue muscles, other anatomical evidence suggests that the larynx in
Neanderthals was placed high in the throat. A highly placed larynx limits an animal’s ability to produce many
sounds, such as vowels. In humans, the larynx is positioned further down into the throat. The Neanderthal has
been determined to have possessed the gene FOXP2, which is linked to the ability to understand complex
language, but the verdict is still out as to whether they were able to produce complex language. It is believed by
some researchers that the ability to produce complex speech gave H. sapiens a significant edge over the
Neanderthal.

Other skull characteristics of the Neanderthal include an occipital bun at the back of the skull (as also seen in
H. antecessor and H. erectus), large brow ridges (which are not solid bone and create an air cavity), a large
nasal cavity, and incisors that show a rounded pattern of wear, especially in older individuals. Their large front
teeth typically show excessive wear. Chipping and pitting on the incisors are believed to have been caused by
chewing on leather. The postcranial bones show that they had a broad scapula, which indicates that their
rotator cuff muscles were well developed. They possessed a robust humerus with a massive head and the
ability to rotate their arms, which suggests they were capable of throwing projectiles and using spears.

Some of the best-known Neanderthal specimens come from a place called Shanidar Cave in Iraq. Within this
cave, various skeletal remains of eight individual Neanderthals were found. These remains are identified as
Shanidar 1–9, which were discovered between 1957 and 1961, and Shanidar 10, which was discovered in
2006. Nearly all the skeletal remains show some evidence of trauma, suggesting that hunting was risky
business. At various Neanderthal sites it has been observed that men and women exhibit similar cranial
injuries, suggesting that women might have also engaged in hunting activities. However, the number of
injuries in women were significantly fewer than those found in men (Beier et al. 2008). In a comparative study,
it was established that during the Upper Paleolithic, modern H. sapiens sustained similar injuries as the
Neanderthal, but interestingly, these injuries were less likely to result in death (Beier et al. 2008).

Shanidar 3 features a 40-to-50-year-old Neanderthal man who suffered a rib injury, potentially as the result of
an encounter with an animal, and suggests healing as a result of care from others. Shanidar 1, called the “Old
Man” (30–45 years old was old in Neanderthal terms), had multiple traumas to his body, one of which resulted
in blindness in one eye. He was also missing the lower part of his right arm and hand, which suggests the
earliest amputation on record. Although he did heal from this amputation, it may have left him paralyzed on
the right side of his body. He also had no teeth. It is believed he was kept alive by taking food that had been
chewed by others for him. There is evidence of many of these individuals healing from their injuries, which
suggests that compassion and a sense of social responsibility for disabled members of the community existed.
160 5 • The Genus Homo and the Emergence of Us

FIGURE 5.22 This cave is the site of the Shanidar 4 Neanderthal flower burial site. Evidence found here and at
other sites indicates that the Neanderthal practiced intentional burials of their dead. (credit: “Shanidar Cave, Iraqi
Kurdistan” by Sammy Six/flickr, CC BY 2.0)

The Flower Burial Hypothesis


The remains found at Shanidar 4 in Iraq suggest that the Neanderthal practiced intentional burials, or
deliberate placing of the dead in a ritualistic manner. At Shanidar 4, the individual is placed on his left side
with his legs drawn up in a flexed position. Pollen analysis of the soil surrounding the corpse suggests that
spring flowers had been placed in the grave, possibly indicating that the Neanderthal had a belief in an afterlife
and established mortuary practices. However, there has been a lot of debate as to whether there is sufficient
evidence to conclude that that the pollen found at some of the Neanderthal sites was a result of ritualistic
placement of flowers. Opposing hypotheses propose that the pollen was brought into the cave and deposited by
burrowing rodents (Sommer 1999). In spite of these counterclaims, the consensus supports the theory that the
Neanderthal did practice intentional burials. This is largely based on evidence such as the careful placement
of bodies in specially dug shallow pits. Recent research at both Shanidar Cave and other sites now support the
claim that the Neanderthal did practice ritual and intentional burial.

FIGURE 5.23 This reconstruction of a Neanderthal grave is housed in the Israel Museum in Jerusalem. (credit:
“Neanderthal Burial, Cast” by Gary Todd/flickr, Public Domain)

Neanderthal Creativity and Material Culture


Neanderthals have been labeled, perhaps unjustly, as a species with a limited ability to communicate in
symbolic or abstract forms. Until recently, the Neanderthal had been assumed to lack the cognitive skills

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 161

associated with the practice of ritual and art. However, cave paintings discovered in Spain in 2012 by Alistair
Pike, an archaeologist at the University of Southampton, UK, challenge that assumption. These paintings,
which have been dated to around 65,000 years ago, before the arrival of H. sapiens in the region, have been
determined to be the creative works of the Neanderthal and are currently considered the oldest cave art ever
found. This discovery may change what people have previously thought about Neanderthal cognition and their
ability to express symbolic thought. It should be acknowledged that the ability to depict the world evident in
these paintings does not compare with that in the artwork from H. sapiens sites like Chauvet and Lascaux in
France (to be discussed later in this chapter).

Neanderthals created more technologically advanced tools than those produced by H. erectus and seen in the
Acheulian tool industry. The tool industry associated with the Neanderthal hominins is called the Mousterian
tool industry or the Middle Paleolithic tool industry. Archeological sites that date to the Neanderthal period
are dominated by flake tools. This means that the Neanderthal struck flakes from cores and then used the
flakes as their tools instead of the core. This resulted in smaller and sharper tools with increased utility.

What Happened to the Neanderthal? What Gave Modern Humans the Edge?
The Neanderthal went extinct around 35,000 to 50,000 years ago. There have been various hypotheses as to
what caused this, many connected to the fact that Neanderthal coexisted with H. sapiens in regions of Europe
and Asia for an estimated 2,600–5,400 years. These hypotheses include an inability to adapt to a changing
climate and colder temperatures, the spread of disease, competition for food with H. sapiens, and even
aggressive takeover by the H. sapiens, who may have been better able to adapt to environmental changes due
to more complex technology and language skills. Another theory points to evidence that the Neanderthal
tended to live in small, scattered groups with limited genetic diversity and low birth rates, which potentially
impacted the ability of the Neanderthal to be competitive. A low gene pool can result from reduced birth rates
and low survival rates of young children. New genetic evidence shows that the Neanderthal were genetically
less diverse and more isolated than H. sapiens. And then some argue that the Neanderthal didn’t go extinct at
all because some people still have Neanderthal genes in them.

Are You a Neanderthal?


Recent genetic evidence indicates that human-Neanderthal interbreeding (https://openstax.org/r/the-scientist)
was happening as far back as 125,000 years ago. From one Neanderthal toe bone found in the Denisova cave in
Siberia Russia, the Max Planck institute has been able to produce a whole genome which revealed evidence of
inbreeding amongst the Neanderthal, along with interbreeding with their cousins the Denisovans (discussed
further in next section), as well as a mystery yet to be identified species, as well as Homo sapiens (Pennisi, E.,
2013). The genetic evidence is most prominent in people of East Asian descent, accounting for between 2.3
percent and 2.6 percent of their DNA. Various mutations and diseases are linked to this Neanderthal DNA,
including diabetes, addictions, depression, allergies, and Crohn’s disease. One study suggests that Neanderthal
genes gave people some level of protection from getting a severe case of COVID-19 (Huber, J., 2018), although a
later study (Zeberg and Pääbo 2020) proposes that Neanderthal genes may have increased the risk of
respiratory failure as a result of the COVID-19 virus. Such differences may have to do with different genetic
clusters in Neanderthal populations in different geographical regions (Mortazavi et al. 2021). Neanderthal
genes are believed to have provided immunity to some viruses that H. sapiens, arriving from Africa, would not
have had time to build up an immunity against. On the reverse side, H. sapiens may have brought diseases
from Africa that the Neanderthal did not have resistance to, possibly playing a role in their extinction. As Janet
Kelso, a computational biologist at the Max Plank Institute for Evolutionary Anthropology, states, “Viral
challenges, bacterial challenges are among the strongest selective forces out there” (Akst, 2019).

The Denisovans
The Denisovans, like Homo naledi, are archaic Homo. There are not a lot of specimens—just one finger bone,
three teeth, some long bone (https://openstax.org/r/Long_bone) fragments, a partial jawbone, and a parietal
bone (https://openstax.org/r/Parietal_bone) skull fragment. Because of this lack of evidence, very little is
known of their anatomical features. Some of the specimens come from Denisova Cave in Siberia, Russia, and
are dated to between 500,000 and 30,000 years ago. These dates are arrived at based on the few fossils that
exist, inferences made from genetic studies, and sediment analysis. More recently another specimen was
found on the Tibetan plateau. In 1980 a jaw and two teeth were uncovered in the Baishiya Karst Cave by a
162 5 • The Genus Homo and the Emergence of Us

monk, but it wasn’t until 2010 that scientists were able to study the jaw. Dating placed the specimen at
approximately 160,000 years ago. Protein analysis determined the jaw to be of Denisovan origin and from a
member of a population who were most likely well adapted to living in high altitudes (Chen et al. 2019).

Because so few bones have been found, most understanding of this species comes from genetic analyses.
According to nuclear DNA studies, Denisovans and Neanderthals were more closely related to each other than
they were to modern humans. DNA evidence suggests that the Denisovans interbred with modern humans and
with local Neanderthal populations over multiple time periods. Tracing the male Y chromosome, one study
indicated that interbreeding between early humans and Neanderthals actually replaced the ancient Denisovan
Y chromosome once found in Neanderthals. The time of divergence of the Denisovan is estimated to be around
700,000 years ago, with modern humans diverging from the Neanderthal around 370,000 years ago (Petr et al.
2020). H. heidelbergensis is typically considered to have been the direct ancestor of both Denisovans and
Neanderthals, and sometimes also of modern humans.

One specimen is a first-generation hybrid, Denisova 11—nicknamed “Denny (https://openstax.org/r/


Denny_hybrid_hominin)”—that had a Denisovan father and a Neanderthal mother (Slon et al. 2018). Denisova
11 was found in Denisova Cave in Russia and provides evidence that Late Pleistocene Homo species interbred
when the groups met. Comparison of the DNA of these three groups suggest that most modern-day Europeans
and Asians inherited about 1–4 percent of their DNA from Neanderthals, with no Denisovan ancestry in
Europe and 0.1 percent in China. The genetics found in Tibetans, Melanesians, and Indigenous Australian are
currently being challenged; originally, they were thought to be about 3–5 percent Denisovan and 2.74 percent
Neanderthal. Statistical geneticist Ryan Bohlender and his team have investigated the percentages of extinct
hominin DNA in modern humans. They concluded that Neanderthals and Denisovans are not the whole story
and that there could be a third group yet unknown contributing to the Pacific Islander genome (Rogers,
Bohlender, and Huff 2017). Statistical and genetic evidence can serve as indicators of the existence of a group
for which no fossils have yet been found. These are referred to as ghost populations. For example, there are
indications that 2–19 percent of the DNA of four West African populations may have come from an unknown
archaic hominin that split from the ancestor of humans and Neanderthals between 360 KYA and 1.02 MYA
(Durvasula and Sankararaman 2020). The hypothesis of a third lineage in the genus Homo appears to have
received further confirmation with a discovery in China.

New Homo Genus Discovery Homo longi, or Dragon Man


Recently a new archaic Homo fossil surfaced in Harbin, China, dated to about 146,000 years ago (Ji et al. 2021).
Given the name H. longi, it has also been called “Dragon man (https://openstax.org/r/dragonman)” as its
origins were determined to be in the province of the Black Dragon River. The fossil (referred to as the Harbin
cranium) was donated to the Hebei GEO University museum after being hidden away in a well in the 1930s
during the construction of a railway bridge. The verdict is still out as to whether H. longi represents a lineage of
the Denisovans or a new species, but it is clear it was robust and able to adapt to one of the coldest regions of
China. It had a large brain, thick brow ridges, and fairly large teeth, similar to what is found in the Denisovans.

Regional Evolutionary Adaptations: Homo floresiensis


The Hobbit of Flores
Homo floresiensis, also known as “the Hobbit” or “Flores Man,” was discovered on the island of Flores in
Indonesia in 2003. The species has been dated to approximately 100,000–60,000 years ago. What was
surprising about this species is its size. An adult individual stood about 3 feet 7 inches tall. Liang Bua, the cave
where H. floresiensis was found, shows evidence of the use of fire for cooking and contains bones with
cutmarks. Since the initial discovery, partial skeletons of nine individuals have been found.

H. floresiensis, like the earlier hominins, did not possess a chin, and its leg bones are thicker than those of
modern humans. They had flat feet that were relatively long in comparison to the rest of their bodies. As a
result of these anatomical differences, it is believed that their bipedalism was quite different from that of
modern humans, with a high stepping gait and slower walking speed. H. floresiensis also had substantially
more mobility in the elbow joint, which suggests that they were tree climbers.

Their small brain size is not believed to have affected their intelligence. This challenges the view that larger

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 163

cranial capacity equals higher cognitive skills. Although H. floresiensis has a brain size of just 380 cc, equal to
the size of an orange, evidence indicates that they made tools, used fire, and hunted very much like H. erectus.
The brain of H. floresiensis does contain a Brodmann area, which is associated with cognitive abilities, that is
the same size as that found in modern humans.

Some have suggested that H. floresiensis is a sister species of Homo habilis that branched off before or shortly
after the evolution of Homo habilis. Other hypotheses suggest that they were the descendants of H. erectus who
became stranded on the island after arriving via water, possibly on bamboo rafts.

Another Homo species similar in size to H. floresiensis was H. luzonensis, found on the island of Luzon in the
Philippines and dated to at least 50,000–67,000 years ago. H. luzonensis displays a hybrid of australopithecine
traits (including curved hands and feet) and Homo characteristics, yet lived alongside modern H. sapiens.
Clearly the genus Homo is more diverse and complex than was originally thought, especially within the special
evolutionary pressures of island environments.

FIGURE 5.24 This H. floresiensis skull is on display at the Naturmuseum Senckenberg, a Natural History Museum
in Germany. (credit: Daderot/Wikimedia Commons, Public Domain)

Island Dwarfism as an Evolutionary Explanation


Numerous hypotheses have been proposed to account for the small brain size found in both H. floresiensis and
H. luzonensis. One initial theory was that H. floresiensis had microcephaly, which is a genetic condition
creating an abnormally small head. This was discounted as an explanation once additional specimens were
found exhibiting the same size. Perhaps the most convincing explanation is an evolutionary theory called
island dwarfism, which notes that the evolutionary pressures on islands can be very different from those
found on the mainland. Island dwarfism posits that mainland small animal species that colonize islands might
evolve larger bodies if the island does not contain key predators. On the other hand, larger species may
become smaller due to more limited resources in an island environment. According to the island dwarfism
hypothesis, H. erectus made its way to Flores, where its descendants became isolated and grew progressively
smaller to make the most of limited resources in the island environment. This theory is supported by the fact
that there are unique sizes displayed by other animals found with H. floresiensis, including a dwarf species of
primitive elephant called a Stegodon. As H. floresiensis’s body shrank, its brain may have undergone
“neurological reorganization” to fit a smaller cranial space while maintaining its brain-to-body ratio. The only
164 5 • The Genus Homo and the Emergence of Us

potential large predator that may have been a threat to H. floresiensis was the Komodo dragon, which ate most
of the large mammals on the island. Nevertheless, predation pressures for the little people were likely quite
low—that is, until H. sapiens arrived.

The Emergence of Us: Homo sapiens


Modern H. sapiens first appeared about 200,000 years ago in Africa. Anthropologists generally classify these
people as “anatomically modern H. sapiens,” which is a way of noting that while their bodies are the same as
modern humans, they had not yet developed the cultural traditions, symbolic behaviors, and technologies that
are seen among later H. sapiens, including people of today. Probably the most defining feature of anatomically
modern H. sapiens is their chin. Modern H. sapiens is the first hominin to exhibit a projecting chin. One of the
most common explanations for this anatomical feature is that the chin evolved in response to human speech
and protects the jaw against stresses produced by the contraction of certain tongue muscles.

Sometime around 40,000 years ago there was an abrupt change in tool technology, subsistence patterns, and
symbolic expression among H. sapiens. These changes seem to have occurred almost simultaneously in
Africa, Asia, Europe, and Australia. While there is evidence of some creative artistic activity in earlier groups
like the Neanderthal, they were not on the same scale as that seen during the Upper Paleolithic, which is also
referred to as “the human revolution.” The level of cultural changes associated with this period has been
compared to the level of change that occurred during the Industrial Revolution of the 19th century.

Among these changes, H. sapiens began assembling a much more elaborate tool kit by constructing tools from
a wider variety of materials including antler, ivory, and bone. During the Upper Paleolithic, humans shifted
from the manufacture of round flakes to the manufacture of blade tools. This construction method is known as
the blade tool industry. Blades are stone flakes that look like a modern knife blades—they are long, thin, and
flat, and they have a sharp edge. They have a much longer cutting edge than flakes do and are thus more
efficient than older technologies. The prepared-core technique of the Mousterian that provided pre-shaped
flakes was refined and extended to create pre-shaped blades.

FIGURE 5.25 This Upper Paleolithic burin tool has a much longer cutting edge than anything that came before it
and was much more efficient than previous technologies. (credit: “Large Knife Upper Paleolithic or later
35000-3900 BCE Africa” by Mary Harrsch/flickr, CC BY 2.0)

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 165

Over the 23,000 years of the Upper Paleolithic, there were many distinctive tool industries within the larger
category of the blade tool industry, including the Aurignacian, Gravettian, Solutrean, and Magdalenian. The
most significant tool during the Upper Paleolithic was the burin. The burin is a narrow-bladed flint capable of
scraping narrow grooves in bone. Scraping two parallel grooves would allow a sliver of bone to be detached as
stock for a needle, pin, or awl.

The Gravettian tool industry lasted from approximately 33,000 to 22,000 years ago. During this tool industry,
there are many instances of animal remains being used for both decorative and traditional tool purposes. For
example, the teeth of arctic foxes were used for decoration, while their arm bones were used as awls and barbs.
Some animal bones such as mammoth tusks and bones were used to not only create tools, but also to make art,
as seen in the Lion figurine in Figure 5.26. This figurine could be the earliest example of a figure having both
human and animal characteristics, a form often associated with shamans or priests. Some have proposed that
the “lion man” is actually a woman due to the lack of a lion mane.

FIGURE 5.26 An ancient figurine of a lion sculpted from a mammoth’s tusk. This figure was discovered in a German
cave in 1939 and dated to around 40,000 years ago, making it one of the oldest figurative sculptures yet discovered
and the earliest example of an animal-shaped figurine. (credit: JDuckeck/Wikimedia Commons, Public Domain)

The Solutrean tool industry utilized tool-making techniques not seen before. It produced finely worked bifacial
points made with lithic reduction percussion rather than flint knapping. Lithic reduction is the process of
fashioning stones or rocks into tools or weapons by removing some parts. The lithic core, such as a partially
formed tool or naturally formed rock, is held in one hand and struck with a hammer or percussor with the
other hand. As flakes are detached, the original mass of stone or lithic core is reduced.

In addition to stone tool innovations, the Solutrean is characterized by the appearance of the atlatl
(https://openstax.org/r/throwing-arrow), or spear thrower. An atlatl is a long stick used to propel a spear or
dart. Functioning as an extension of the arm, this stick of wood or antler added kinetic energy, and therefore
range, to a short spear tipped with flint or bone. The earliest archeological evidence for this tool innovation
comes from France, where a 17,500-year-old atlatl was found constructed out of reindeer antler. It is believed
that the atlatl was used by humans to hunt large fauna.
166 5 • The Genus Homo and the Emergence of Us

FIGURE 5.27 Contemporary man using an atlatl, a tool for launching a spear or a dart that is at least 17,500 years
old. (credit: “Atlatl throwing demonstration” by Hannah Schwalbe/NPS/flickr, Public Domain)

By 17,000 years ago, the Solutrean tool industry was replaced by a new tool industry known as the
Magdalenian tool industry. During this period, bone and ivory continue to be used, as well as stone. Unlike
Mousterian tools, Solutrean tools are made not only from nearby rocks, but also from rocks that have been
transported over relatively long distances. Keep in mind that this required not only transporting the selected
rocks, but also finding and extracting them.

The Gravettian tool industry is best known for carved Venus figurines portraying a woman, typically made
from ivory or limestone. Most figurines have small heads, wide hips, and large breasts. Most researchers
believe that they served a ritual or symbolic function. Some have suggested that they represent an expression
of health and fertility.

FIGURE 5.28 Venus of Hohle Fels figurine. This figurine is considered to be the earliest known depiction of a
human being in prehistoric art. (credit: Anagoria/Wikimedia Commons, CC BY 3.0)

Access for free at openstax.org


5.3 • The Emergence of Us: The Archaic Homo 167

During the Upper Paleolithic, H. sapiens created a great deal of cave art. More than 350 cave painting sites
have been discovered, the majority located in France and Spain. Cave art seems to have been created
continually from 40,000 to 10,000 years ago and then disappeared around 10,000 years ago, likely due to
climate change. As temperatures increased, underground shelters were gradually replaced by surface
settlements. The most well-known cave sites in France are the Chauvet (https://openstax.org/r/chauvet-
archeologie-culture) (32,400 years ago) (Figure 5.29) and Lascaux Caves (17,000 years ago). The art in both
caves features common subjects, such as bison, horses, and deer, as well as tracings of human hands. Most of
the animals depicted were commonly hunted but were not always found with associated deposits of bones. The
cave art produced during the Upper Paleolithic show a level of sophistication and even sacredness not seen
previously in human history.

FIGURE 5.29 These drawings of lions from the Chauvet Cave in France are dated to 32,400 years ago. (credit:
HTO/Wikimedia Commons, Public Domain)

FIGURE 5.30 Handprints found in the Cuevas de las Manos upon Río Pinturas, near the town of Perito Argentina.
Hand stencils on cave walls have been found in many locations around the world. (credit: “SantaCruz-CuevaManos-
P2210651b” by Golan Levin/flickr, CC BY 2.0)
168 5 • The Genus Homo and the Emergence of Us

Cave paintings were made with natural pigments created by mixing ground-up elements, such as dirt, red
ochre, hematite, manganese oxide, and animal blood, with animal fat and saliva. Paint was applied using twigs
formed into brushes and blow pipes made from bird bones, through which paint was sprayed onto the cave
wall. Hand stencils on cave walls can be found in many locations around the world including Africa, Argentina,
Europe, and Australia. Anthropologist Dean Snow (2013) conducted research at eight cave sites in France and
Spain to determine who the artists might be. Based on calculated measurements of the handprints, he
concluded that 75 percent of the ochre stenciled handprints in the Paleolithic caves were made by women.

5.4 Tracking Genomes: Our Human Story Unfolds


LEARNING OBJECTIVES

By the end of this section, you will be able to:


• Describe how mtDNA sheds light on early human migrations and explain the Out of Africa model.
• Explain how studying the genomes and coevolution of lice can fill current gaps in the human fossil record.
• Describe the origin of human variation from an evolutionary perspective.

Mitochondrial Eve
Begun in 1990 and concluding in 2003, the Human Genome Project was an ambitious international effort that
sequenced about 99 percent of the human genome with an accuracy of 99.99 percent. Genetics has thus far
largely confirmed the Out of Africa theory, which proposes that early humans left Africa around 100,000
years ago and migrated to diverse areas of the world. When early humans left Africa and moved into Europe,
they not only lived alongside but also interbred with non-African species such as the Neanderthal, who were
already inhabiting the region.

Molecular anthropologists have an interest in determining when living human populations began diverging
from one another. This has been difficult to do using nuclear DNA because it mutates much too slowly for
measurable accumulations to occur in 200,000 years. Many of the genetic studies that have been conducted
are thus based on genetic material carried in the mitochondria (mtDNA), which are passed on maternally.
There is no recombination in mtDNA, so unless the mitochondria carries a novel mutation, a child has exactly
the same mitochondrial genes as its female genetic contributor (which may be its mother, egg donor, or
someone in a similar genetic relationship). The mitochondria of every living person is a copy, modified only by
rare mutations, of the mitochondria passed down via matrilineal descent from a population in our ancient
past. This population is referred to as Mitochondrial Eve or mtMRCA (mitochondrial most recent common
ancestor), believed to have lived in southern Africa 100,000–200,000 years ago.

As discussed in Chapter 4, the longer ago two populations share a common ancestor, the more time there is for
mutations to occur and for adaptations and change to take place. Although genetic variation is small among
the world’s human populations, it is greatest in Africa. This indicates that the human populations in Africa
have the longest established genetic lineage. While multiple hypotheses exist as to human origins and new
evidence could change the current views, the consensus is an Out of Africa model traced back to the
matrilineal descent of a population living in Africa about 200,000 years ago.

How the Genome of Lice Can Fill in the Gaps


While perhaps not a pleasant thought, lice have long been a part of human history. Studying the coevolution
relationship between humans and lice has shed much light on human story. Dr. David Reed, the Curator of
Mammals and Associate Director of Research and Collections at the University of Florida Museum, has been
studying the coevolution of humans and lice, an area of research that has developed only within the last 20
years. Reed’s groundbreaking research has the potential to fill in some big gaps in humans’ rather sketchy
fossil record and provides important data that might have applications in medicine and biology. Two questions
that this research has already begun to ask are when did we become less hairy and when did we start wearing
clothes.

Access for free at openstax.org


5.4 • Tracking Genomes: Our Human Story Unfolds 169

FIGURE 5.31 There are three types of lice associated with humans: (a) crab or pubic louse; (b) body louse; (c))
head louse. The coevolution of humans and lice is a developing area of research. (credit: (a) Noizyboy1961/
Wikimedia Commons, CC BY 4.0; (b) Janice Harney Carr, Centers for Disease Control and Prevention/Wikimedia
Commons, Public Domain; (c) Dr. Dennis D. Juranek, Centers for Disease Control and Prevention/Wikimedia
Commons, Public Domain)

Figure 5.31 shows three types of lice associated with humans: the head louse (Pediculus humanus capitis), the
body louse (Pediculus humanus corporis), and the crab louse or pubic louse (Pthirus pubis). Body lice infest
clothing and lay their eggs on fibers in the fabric. Head and pubic lice infest hair, laying their eggs at the base of
hair fibers. The human head and body lice (genus Pediculus) share a common ancestor with chimpanzee lice,
while crab lice (genus Pthirus) share a common ancestor with gorilla lice. By tracking louse variations,
scientists have been able to determine when the head louse and pubic louse diverged, enabling estimates as to
when we lost our extra hair and when we started to wear clothes. It is interesting to note that the divergence of
the genus Pediculus (head and body lice) correlates with the divergence of the human lineage from
chimpanzees about six million years ago. Research on lice also provides further support for the Out of Africa
model of human migration. Reed has observed that the genome of African lice shows a higher degree of genetic
diversity than that of lice found elsewhere in the world, supporting the hypothesis that both humans and lice
existed in Africa first.

Many hypotheses about what may have triggered the loss of hair in humans point to thermoregulation, the
need to control body temperature in extreme conditions. Living in the heat of the savanna, humans needed a
cooling mechanism to enable them to be better hunters. Other evidence of adaptation to the heat includes the
appearance of sweat glands, which are more numerous in humans than in other primates. Another theory
about the cause of the loss of hair among humans suggests that it was an adaptation to control parasites on the
body. Did people immediately throw on clothes after losing all of that extra body hair? Reed’s research suggests
that the wearing of clothes was not something that happened quickly. Humans lost body hair about a million
years ago and didn’t start wearing clothes until around 170,000 to 190,000 years ago. That’s about 830,000
years living in their birthday suits! When humans began to wear clothes, the body louse adapted structures
that enabled them to attach to clothes instead of hair.
170 5 • The Genus Homo and the Emergence of Us

FIGURE 5.32 Humans lost most of their body hair about a million years ago. (credit: “Neanderthal” by Eden, Janine
and Jim/flickr, CC BY 2.0)

PROFILES IN ANTHROPOLOGY

Molly Selba

FIGURE 5.33 Molly Selba (holding skull) leading a study session. (credit: Molly Selba, Public Domain)

Personal History: From the time Molly was young, she knew she wanted to be an anthropologist.

She took an archaeology class at the local community college when she was a high school student and went to
field school over the summer. In college, she completed a double major in archaeology and anthropology, with
a minor in Museums and Society. She later gained experience working with different museum collections and
held internships at the Baltimore City Medical Examiner’s Office and the Smithsonian Museum of Natural
History. After completing her undergraduate degrees, she knew that she wanted to pursue anthropology as a
full-time career and began working towards her master’s and doctorate in biological anthropology.

Access for free at openstax.org


5.4 • Tracking Genomes: Our Human Story Unfolds 171

Area of Anthropology: For Molly the most interesting thing about biological anthropology is the information
that bones can tell us. Initially she was interested in what the history of disease could tell us about the lives of
people in the past, but as she worked with biological anthropologists, her focus shifted to understanding how
evolution can impact the shape of different bones.

She received her undergraduate degree from Johns Hopkins University in Baltimore, Maryland, and her
Master’s degree from the University of Florida, where she is currently a PhD candidate. Her research interests
include comparative anatomy, cranial morphology, and anatomical sciences education. She is most interested
in how cranial morphology varies within and between species and how it is impacted by factors such as
evolution and selective breeding practices. Her earlier research focused on the differences in cranial
morphology in dogs created by artificial selection for facial reduction. Her dissertation research currently
focuses on a comparative study of facial reduction across bats, primates, and dogs.

Accomplishments in the Field: For Molly her most important accomplishment in the field of anthropology has
been in education and outreach. Throughout her time in graduate school, she was involved in school visits,
working with teachers to facilitate the inclusion of human evolution into existing science curricula. She has
specifically focused on helping educators find teaching materials that are culturally inclusive and responsive.
She has led multiple professional development workshops for teachers on the same topic and has visited over
two dozen classrooms and interacted with over 1,200 students in the last four years. Making science accessible
to K-12 educators is an extremely important part of being a researcher, and she believes everyone in academia
should strive to be effective science communicators.

“Studying biological anthropology helps us better understand our origin story as a species. It helps us
recognize why our anatomy is the way that it is, how morphological changes over time can take place, and why
we have such a diversity of life on earth. Just being able to recognize and identify our anatomy is only half the
challenge—more important is our understanding why various traits are adaptive, how structure relates to
function, or why leftover anatomical traits still persist in our body to this day.”

Natural Selection and Human Variation: Are Humans Still Evolving?


Human variability is attributed to a combination of environmental and genetic factors, including social status,
ethnicity, age, nutrition, quality of life, access to healthcare, work and occupation, etc. As mentioned in
Chapter 1, anthropology contributes many insights into both the social construct of race and the impacts racial
categories have on people’s lives. The focus in this chapter is the role of natural selection in human variation.

A number of changes are associated with the Neolithic era and the rise of agriculture around 10,000 to 8,000
years ago. Many have noted that changes during this time period did not have positive effects on human and
environmental health. The evolutionary mismatch hypothesis proposes that our bodies are best suited to the
environments we have spent much of our evolutionary history in, which are very different from the
environments we inhabit today (Li, van Vugt, and Colarelli 2018).

Humans evolved for one million years as hunter-gatherers. Today, human bodies are still trying to adapt to the
largely grain-based diet brought about by agriculture, a diet characterized by less diversity and lower levels of
nutrition than that of a typical hunter-gatherer. Incomplete adaptation to this change has made people
susceptible to a number of diseases and nutritional deficiencies. Lactose intolerance is a prime example. The
domestication of cattle and the drinking of cow’s milk began during the agricultural age, not very long ago in
evolutionary history. Currently 65 percent of humans are unable to digest cow’s milk. Dental caries (cavities)
are another problem linked to the change in diet associated with agriculture. The grain-based and high-sugar
diets associated with agriculture are very different from the diet of hunter-gatherers. Neither our bodies nor
the bacteria in our mouths have had time to fully adapt to this change.

Another adaptation that took place during the Neolithic era is related to variation in skin pigmentation.
Humans who left Africa and settled in Europe about 40,000 years most likely had dark skin with high levels of
melanin, which provides protection against ultraviolet radiation New data confirms that about 8,500 years
ago, early hunter-gatherers in Spain, Luxembourg, and Hungary also had darker skin. Skin pigmentation is an
adaptation to ultraviolet radiation, with different tones offering different advantages, depending on one’s
172 5 • The Genus Homo and the Emergence of Us

distance from the equator. As humans migrated to the Northern Hemisphere, they were exposed to less
ultraviolet radiation, which also meant less absorption of the Vitamin D needed for strong bones and other
important immune functions. In order to compensate for this loss and to allow for greater exposure to
ultraviolet radiation, skin pigmentation became lighter.

Another example of human variation as a result of adaptation to the environment can be seen in Indigenous
populations in the Andes, Tibet, and the Ethiopian highlands. Each of these three groups faces the same
environmental challenge, living in a low-oxygen environment, and they have responded with unique
adaptations. Tibetans compensate for low oxygen levels by taking more breaths per minute than people who
live at sea level. Those living at high altitudes in the Andes have been found to have higher concentrations of
hemoglobin in their blood than other people. Ethiopians living at altitudes of 9,800 to 11,580 feet have neither
of these adaptations. The explanation as to how the Ethiopian highlanders thrive in their environment is still a
mystery.

FIGURE 5.34 A valley in the Andes near Ollantaytambo, Peru. Indigenous peoples living at high altitudes in the
Andes have been found to have higher concentrations of hemoglobin in their blood than other people. (credit:
“Snows of the Andes” by David Stanley/flickr, CC BY 2.0)

This chapter has explored just some of the immense biological and cultural diversity of the genus Homo. This
diversity has emerged in response to highly complex and variable environments connected to factors such as
exposure to UV radiation, low oxygen levels at high altitude, changes in diet as a result of hunting or
agricultural practices, geographic isolation in island populations, and climate variability and temperature. The
genus Homo has proven to be resilient and adaptive in response to whatever environment or challenge it has
faced. Variation is the key to survival. While scientists recognize that biological and cultural variation has
greatly contributed to our human evolution, the human species is now facing a moment in which we must
contemplate a difficult question: To what extent has our success as a species jeopardized the survival of other
species and the health of the planet we all call home?

MINI-FIELDWORK ACTIVITY

Identify the Fossil

Imagine that you have just discovered a hominin fossil with some of the characteristics listed below. Write each
of the characteristics on a card and shuffle them together. Then, working in a group, decide which
characteristics belong in the Homo group and which belong in the Australopithecus group. What scientific
name (genus/species) would you give it, and what criteria did you use? (Note: This is an actual hominin fossil!)

Access for free at openstax.org


5.4 • Tracking Genomes: Our Human Story Unfolds 173

• Brain similar in shape and structure to modern human brains


• Hands suited for tool use
• Small jaws and teeth
• Third molar larger than other molars (found in australopithecines and some early Homo species)
• Skull shaped more like H. erectus or H. habilis
• A sagittal keel (as seen in H. erectus)—a small raised ridge on top of the skull
• Bipedal and walked with a human gait
• Humanlike feet with arches and ankles
• Flaring blades of the pelvis (primitive)
• Broad rib cage
• Lower part of pelvis like modern humans’
• Small braincase (EQ 4.5)
• Skull shows prognathism (protruding face)
• Primitive shoulder position suggests suitability for climbing and swinging
• Curved fingers (What would that suggest?)

Additional Resources
Visual timelines and maps

The Atlas of Human Evolution has put together a user friendly interactive map (https://openstax.org/r/
atlasofhumanevolution) on the development of Homo sapiens.

The Smithsonian Institute's Human Evolution Interactive Timeline (https://openstax.org/r/human-evolution-


interactive-timeline) provides an interactive overview of major milestones and species ranges.

Coevolution of lice and humans

David Reed, associate curator of mammals at the Florida Museum of Natural History, offers an explanation of
his research (https://openstax.org/r/liceshowhumans) for a general audience.

The Smithsonian Magazine explores possible causes and benefits (https://openstax.org/r/smithsonianmag) of


the human loss of hair.
174 5 • Key Terms

Key Terms
Acheulean tool industry the production of more island dwarfism mainland small animal species
complex tools, including hand axes, by H. erectus that colonize islands might evolve larger bodies if
from 1.6 million to 200,000 years ago. the island does not contain key predators. On the
archaic Homo the period of time that precedes the other hand, larger species may become smaller
emergence of the earliest early modern humans due to more limited resources in an island
(Homo sapiens) around 300,000 years ago. environment.
biface tools a type of tool characteristic of the lithic reduction the process of fashioning stones
Acheulean tool industry, with both sides worked. or rocks into tools or weapons by removing some
Broca’s area a region in the frontal lobe of the parts.
brain (which includes two Brodmann areas) first melanin substance that determines the color of
found in H. habilis and connected with the skin pigmentation and protects people from
production of speech. ultraviolet radiation. Skin will have higher levels
canine fossa a facial depression above the canine of melanin the closer to the equator one lives.
tooth found in modern humans. Mitochondrial Eve genes traced through
coevolution an interaction between different mitochondrial DNA that represent the female
species that influences each species’ evolution; genetic originator of all humans who lived
the simplest case of this is predator-prey 200,000 years ago in Africa.
relationships. Mousterian tool industry a complex stone tool
cranial capacity the volume of the interior of the technology largely associated with the
cranium or skull, providing an approximate size Neanderthal.
of the brain. mtDNA the DNA located in the mitochondria that
encephalization increased brain size over time. can be passed down unchanged from female
encephalization quotient a measurement defined genetic contributor to child.
as the ratio between brain and body size. mtMRCA “mitochondrial most recent common
endocranial cast an impression taken from the ancestor,” or Mitochondrial Eve, representing the
inside of the cranium (braincase), frequently common ancestor of H. sapiens around 200,000
used by paleoanthropologists to determine the years ago.
shape and approximate size of the brain in occipital bun an anatomical feature seen in the
hominids and other primates. Neanderthal skull that appears in the rear of the
evolutionary mismatch a hypothesis that disease skull.
and nutritional deficiencies result when people’s Oldowan tool Industry the oldest and most
bodies are unable to adapt to an environment that primitive tool industry; production and use are
they have not spent most of their evolutionary largely in association with H. habilis.
history in. Out of Africa theory theory that proposes that
genome the complete set of genes or genetic Homo sapiens developed first in Africa and then
material present in a cell or organism. spread around the world between 100,000 and
ghost population proposed group for which no 200,000 years ago.
fossil evidence has yet been found. postcranial features skeletal material found in the
hafting the process of attaching stone points to a body that is not related to the skull (cranial
handle, which increases a tool’s effectiveness for bones).
hunting. sexual dimorphism differences in physical
handedness the use of a dominant hand, suggests characteristics other than reproductive organs
lateralization of the brain and cognitive that appear between males and females of the
development. same species.
intentional burials evidence of placing the dead in thermoregulation an adaptation that allows the
a specific manner, suggesting ritualistic practice. body to control and regulate body temperature.

Summary
In this chapter we have explored our human journey First on the scene were the australopithecines, who
as a member of the genus Homo, following a trail of were already walking on two feet and paved the way
adaptations and change that ultimately led to us. for the evolutionary changes and cultural

Access for free at openstax.org


5 • Critical Thinking Questions 175

achievements that were to follow. A colder climate impacting people today. From the earliest
with drastic changes in climate were associated with toolmakers to the cave art of the Upper Paleolithic to
an increased reliance on cooked meat, which may the modern computer age, the predominant theme
have contributed to a growing brain. A brain with of human history has always been about change. The
highly developed cognitive skills gave humans the ability to adapt to this change is why humans are
capacity to solve problems and create tools that still here. Humans’ evolutionary story, however, does
enabled better hunting and survival skills. not end with the emergence of the species. Today
Adaptations provided H. ergaster and H. erectus the humans are faced with numerous challenges as they
ability to walk and run longer distances, to more adapt to an increasingly changing environment as a
effectively track and follow game, and to explore result of climate change, loss of habitat, and
nearby continents. decreasing biodiversity. In 2020, Darwin’s theory of
natural selection played out in real time as people
Genetic information provided by mtDNA indicates
began an arms race with a mutating and evolving
that all humans shared common ancestors who lived
COVID virus. Evolutionary change is not something
in Africa 200,000 years ago. Studies of genetics
that happened to people just in the past—it is very
shows examples of coevolution and how even small
much still happening today, and it will continue to
organisms such as lice can shed light on the human
be part of the future.
story. The rise of agriculture created new challenges
for humanity, with evolutionary mismatch still

Critical Thinking Questions


1. What are some of the key anatomical differences 5. Based on current evidence, form a hypothesis as
between Australopithecus and one of the species to what you think caused the extinction of the
in the genus Homo? Neanderthal. What gave modern humans the
2. What criteria would you use to define a species edge?
belonging to the genus Homo? 6. In what ways has genetic research enabled
3. What are some of the limitations in using the modern humans to track their story?
currently known criteria for classifying a species 7. What are some similarities between Homo
under the genus Homo? naledi, Homo floresiensis, and Homo luzonensis,
4. What are some of the explanations or hypotheses and what makes them so unique?
for the increasing brain size (encephalization) 8. In what way are humans still evolving today? Can
seen in the genus Homo? you provide an example?

Bibliography
Ahern, Jim. 2015. “Archaic Homo.” In Basics in Human Evolution, edited by Michael P. Muehlenbein, 163–76.
Cambridge, MA: Academic Press. doi:10.1016/B978-0-12-802652-6.00012-8.

Aiello, L., and P. Wheeler. 1995. “The Expensive-Tissue Hypothesis: The Brain and the Digestive System in
Human and Primate Evolution.” Current Anthropology 36 (2): 199–221. http://www.jstor.org/stable/
2744104.

Akst, Jef. 2019. “Neanderthal DNA in Modern Human Genomes Is Not Silent.” The Scientist, September 1,
2019. https://www.the-scientist.com/features/neanderthal-dna-in-modern-human-genomes-is-not-
silent-66299.

Beier, J., N. Anthes, J. Wahl, and K. Harvati. 2018. “Similar Cranial Trauma Prevalence among Neanderthals
and Upper Paleolithic Modern Humans.” Nature 563:686–90. https://doi.org/10.1038/s41586-018-0696-8.

Boaz, Noel T., Russell L. Ciochon, Qinqi Xu, and Jinyi Liu. 2004. “Mapping and Taphonomic Analysis of the
Homo erectus Loci at Locality 1 Zhoukoudian, China.” Journal of Human Evolution 46 (5): 519–49.
doi:10.1016/j.jhevol.2004.01.007.

Bramble, D., and D. Lieberman. 2004. “Endurance Running and the Evolution of Homo.” Nature 432:345–52.
https://doi.org/10.1038/nature03052.

Chen, F., F. Welker, C. C. Shen, S. E. Bailey, I. Bergmann, S. Davis, H. Xia et al. 2019. “A Late Middle Pleistocene
176 5 • Bibliography

Denisovan Mandible from the Tibetan Plateau.” Nature 569:409–12. https://doi.org/10.1038/


s41586-019-1139-x.

Dirks, Paul HGM, Lee R. Berger, Eric M. Roberts, Jan D. Kramers, John Hawks, Patrick S. Randolph-Quinney,
Marina Elliott et al. 2015. “Geological and Taphonomic Context for the New Hominin Species Homo naledi
from the Dinaledi Chamber, South Africa.” eLife 4:e09561. doi:10.7554/eLife.0956.

Dorey, Fran. 2020. “Homo erectus.” Australian Museum. https://australian.museum/learn/science/human-


evolution/homo-erectus/.

Dorey, Fran. 2019. “Homo naledi.” Australian Museum. https://australian.museum/learn/science/human-


evolution/homo-naledi/.

Dunsworth, Holly M. 2010. “Origin of the Genus Homo.” Evolution: Education and Outreach 3:353–66.
doi:10.1007/s12052-010-0247-8.

Durvasula, Arun, and Sriram Sankararaman. 2020. “Recovering Signals of Ghost Archaic Introgression in
African Populations.” Science Advances 6 (7): eaax5097. doi:10.1126/sciadv.aax5097.

Egeland, Charles P., Manuel Domingues-Rodrigo, Travis R. Pickering, Colin G. Menter, and Jason L. Heaton.
2018. “Hominin Skeletal Part Abundances and Claims of Deliberate Disposal of Corpses in the Middle
Pleistocene.” PNAS 115 (18): 4601–06. https://doi.org/10.1073/pnas.1718678115.

Fuentes, Agustin. 2019. Biological Anthropology: Concepts and Connections. 3rd edition. New York: McGraw-
Hill Education.

Frayer, David & Clarke, Ronald & Fiore, Ivana & Blumenschine, Robert & Pérez-Pérez,FF Alejandro & Martínez
Martínez, Laura & Estebaranz-Sánchez, Ferran & Holloway, Ralph & Bondioli, Luca. (2016). OH-65: The
earliest evidence for right-handedness in the fossil record. Journal of Human Evolution. 100. 10.1016/
j.jhevol.2016.07.002.

Hawks, John, Marina Elliott, Peter Schmid, Steven E. Churchill, Darryl J. de Ruiter, Eric M. Roberts, Hannah
Hilbert-Wolf et al. 2017. “New Fossil Remains of Homo naledi from the Lesedi Chamber, South Africa.” eLife
6:e24232. doi:10.7554/eLife.24232.

Holloway, Ralph L. 2018. The International Encyclopedia of Biological Anthropology. Hoboken, NJ: John Wiley
and Sons.

Huber, J. 2018. “Inherited Neanderthal Genes Protect Us Against Viruses, Study Shows.” Scope, October 5,
2018. https://scopeblog.stanford.edu/2018/10/05/inherited-neanderthal-genes-protect-us-against-viruses-
study-shows.

Jerison, H. J. 1973. Evolution of Brain and Intelligence. New York: Academic Press.

Ji, Qiang, Wensheng Wu, Yannan Ji, Qiang Li, and Xijun Ni. 2021. “Late Middle Pleistocene Harbin Cranium
Represents a New Homo Species.” The Innovation 2 (3): 100132. https://www.cell.com/the-innovation/
fulltext/S2666-6758(21)00057-6.

Leakey, L. S., P. V. Tobias, and J. R. Napier. 1964. “A New Species of the Genus Homo from Olduvai Gorge.”
Nature 202:7–9. doi:10.1038/202007a0.

Li, Norman P., Mark van Vugt, and Stephen M. Colarelli. 2018. “The Evolutionary Mismatch Hypothesis:
Implications for Psychological Science.” Current Directions in Psychological Science 27 (1): 38–44.
https://doi.org/10.1177/0963721417731378.

Lieberman, P. 2007. “The Evolution of Human Speech: Its Anatomical and Neural Bases.” Current
Anthropology, 48 (1): 39-66. doi:10.1086/509092.

Meyer, Matthias, Juan-Luis Arsuaga, Cesare de Filippo, Sarah Nagel, Ayinuer Aximu-Petri, Birgit Nickel,
Ignacio Martínez et al. 2016. “Nuclear DNA Sequences from the Middle Pleistocene Sima de los Huesos
Hominins.” Nature 531: 504–07. https://doi.org/10.1038/nature17405.

Access for free at openstax.org


5 • Bibliography 177

Monfils, A., J. Allen, B. W. Goodner, and D. Linton. 2020. “Louse and Human Coevolution.” Biodiversity Literacy
in Undergraduate Education, QUBES Educational Resources. doi:10.25334/JJBH-SG27.

Mortazavi, S., K. Kaveh-Ahangar, S. Mortazavi, D. Firoozi, and M. Haghani. 2021. “How Our Neanderthal Genes
Affect the COVID-19 Mortality: Iran and Mongolia, Two Countries with the Same SARS-CoV-2 Mutation
Cluster but Different Mortality Rates.” Journal of Biomedical Physics and Engineering 11 (1): 109–14.
https://doi.org/10.31661/jbpe.v0i0.2010-1218.

Nielsen, R., J. Akey, M. Jakobsson, J. Pritchard, S. Tishkoff, and E. Willersley. 2017. “Tracing the Peopling of the
World through Genomics.” Nature 541: 302–10. https://doi.org/10.1038/nature21347.

Petr, Martin, Mateja Hajdinjak, Qiaomei Fu, Elena Essel, Hélène Rougier, Isabelle Crevecoeur, Patrick Semal et
al. 2020. “The Evolutionary History of Neanderthal and Denisovan Y Chromosomes.” Science 369 (6511):
1653–56. doi: 10.1126/science.abb6460.

Pike, A. W. G., D. L. Hoffmann, M. García-Diez, P. B. Pettitt, J. Alcolea, R. De Balbín, C. González-Sainz, C. De Las


Heras, J. A. Lasheras, R. Montes et al. (2012). “U-Series Dating of Paleolithic Art in 11 Caves in Spain.”
Science 336 (6087): 1409–13 doi:10.1126/science.1219957.

Schaik, Carel P. V., Triki, Zegni, Bshary, R. Heldstab, S.A., (2021) “Farewell to the Encephalization Quotient: A
New Brain Size Measure for Comparative Primate Cognition” Brain Behav Evol 96:1–12 van Schaik/Triki/
Bshary/Heldstab DOI: 10.1159/000517013

Toups, M.A. Kitchen, A., Light, J.E. Reed, D.L (2011). “Origin of clothing lice indicates early clothing use by
anatomically modern humans in Africa” Molecular Biology and Evolution, Volume 28, Issue 1, January
2011, Pages 29–32, https://doi.org/10.1093/molbev/msq234

Read, Dwight, and Sander van der Leeuw. 2008. “Biology Is Only Part of the Story.” Philosophical Transactions
of the Royal Society 363 (1499): 1959–68. doi:10.1098/rstb.2008.0002.

Rogers, A. R., R. J. Bohlender, and C. D. Huff. 2017. “Early History of Neanderthals and Denisovans.” PNAS 114
(37): 9859–63. doi:10.1073/pnas.1706426114.

Roth, G., and U. Dicke. 2012. “Evolution of the Brain and Intelligence in Primates.” Progress in Brain Research
195:413–30. doi:10.1016/B978-0-444-53860-4.00020-9.

Semaw, S., P. Renne, J. W. K. Harris, C. S. Feibel, R. L. Bernor, N. Fesseha, and K. Mowbray. 1997. “2.5-Million-
Year-Old Stone Tools from Gona, Ethiopia.” Nature 385:333–6. https://doi.org/10.1038/385333a0.

Slon, Viviane, Fabrizio Mafessoni, Benjamin Vernot, Cesare de Filippo, Steffi Grote, Bence Viola, Mateja
Hajdinjak et al. 2018. “The Genome of the Offspring of a Neanderthal Mother and a Denisovan Father.
Nature 561:113–6. https://doi.org/10.1038/s41586-018-0455-x.

Snow, Dean. 2013. “Sexual Dimorphism in European Upper Paleolithic Cave Art.” American Antiquity 78 (4):
746–61. doi:10.7183/0002-7316.78.4.746.

Sommer, J. D. 1999. “The Shanidar IV Flower Burial: A Re-evaluation of Neanderthal Burial Ritual.” Cambridge
Archaeological Journal 9 (1): 127–9. https://doi.org/10.1017/S0959774300015249.

“Study Finds Earliest Evidence in Fossil Record for Right-Handedness.” 2016. University of Kansas.
https://news.ku.edu/2016/07/21/study-finds-earliest-evidence-fossil-record-right-handedness.

Susman, R. 1991. “Who Made the Oldowan Tools? Fossil Evidence for Tool Behavior in Plio-Pleistocene
Hominids.” Journal of Anthropological Research, 47 (2), 129–51. http://www.jstor.org/stable/3630322

Xiujie, Wu, Liu Wu, and Christopher J. Norton. 2007. “Endocasts —The Direct Evidence and Recent Advances in
the Study of Human Brain Evolution.” Progress in Natural Science 17 (9).

Zeberg, H., and S. Pääbo. 2020. “The Major Genetic Risk Factor for Severe COVID-19 Is Inherited from
Neanderthals.” Nature 587:610–2. doi:10.1038/s41586-020-2818-3.
178 5 • Bibliography

Access for free at openstax.org


CHAPTER 6
Language and Communication

Figure 6.1 Family members gather at a sweet potato festival in Gushegu in northern Ghana. This highly social
event brought together families, farmers, chiefs, and community members to celebrate the harvest of sweet
potatoes. (credit: Official photographer of the US Embassy in Ghana/USAID in Ghana/Wikimedia Commons, Public
Domain)

CHAPTER OUTLINE
6.1 The Emergence and Development of Language
6.2 Language and the Mind
6.3 Language, Community, and Culture
6.4 Performativity and Ritual
6.5 Language and Power

INTRODUCTION Talk, talk, talk. As human beings, that is what we do all day (and sometimes all night). Even
when we are alone, we might be listening to the radio, watching a video, reading, or texting—all activities that
incorporate language. Language is often considered to be one of the quintessential elements of humanity, key
to our social interactions and cultural development. No other animal does it the way we do. A few apes have
been taught words in sign language, mainly using simple word combinations to ask for particular treats or
desired activities. Is that anything compared to what we do with language?

Consider a situation from the author, Jennifer Hasty’s own fieldwork.

While conducting research in Ghana, I once attended a large family gathering to honor the birth of a
child, an event called an “outdooring.” After everyone had arrived and socialized a bit, a middle-aged
man stood up and took the microphone in his hand to pour libation. Libation is the ritual offering of
180 6 • Language and Communication

drink to the ancestors, welcoming them to the ceremony and asking for their blessings. As he took the
cup in his hand, he surveyed his audience, then stopped short, appearing extremely embarrassed.
Looking down at his feet, he sputtered, “Oh! When the tongue is present, the teeth do not make noise.”

Everyone laughed. It was a proverb I’d heard before, but I had no idea what it meant in this context.
The speaker stepped aside as an even older man rose from a table at the edge of the gathering and
slowly made his way to the microphone. The first speaker had assumed he was the eldest member of
the family present at the gathering, but in fact, his older brother was there. By the rules of seniority, it
was the older brother who should present the libation.

What did the proverb mean in that situation? In most cultures, people do not usually explain proverbs, so the
listener has to piece together the meaning. In this case, the proverb was used metaphorically to compare the
production of words in the mouth and the roles of the people involved in this particular performance of
libation. The nimble tongue is central to human speech, while the teeth play a more fixed and supportive role,
providing surfaces used by the tongue to make certain sounds. Alone, the teeth can only clash against each
other meaninglessly. A tongue is needed to produce speech. Using the proverb, the first speaker was
comparing his elder brother to the tongue—he was more central to the gathering and more proficient in the
production of ceremonial speech such as libation. The younger man assigned himself the role of a tooth, only
able to make noise rather than ceremonial speech.

In humans, language has developed into an extremely complex feature of sociocultural life. Just as the tongue
is central to the production of human speech, language is central to the production of human culture. The
subfield of linguistic anthropology examines the role of language in sociocultural life. Linguistic
anthropologists are interested in how language affects our thinking and our experience of the world around us.
Some explore the different categories of formal and informal speech that people have developed to organize
rituals and ceremonies as well as everyday activities. Others listen carefully to various kinds of conversation,
looking for patterns in the way people interpret and build on one another’s speech acts.

The discipline linguistics is devoted to the study of language. Linguistics is the science of language, including
subfields devoted to speech sounds, word forms, word arrangement, meanings, and practical language use.
One subfield of linguistics, sociolinguistics, examines the social context of language use, such as how language
varies according to age, gender, class, and race. While sociolinguistics and linguistic anthropology share an
interest in the social side of language, linguistic anthropologists tend to focus on language as an aspect of
larger cultural processes. Rather than looking at language as a sole object of study, linguistic anthropology
studies language as one cultural element among many, all interwoven into the sociocultural life of a people.

6.1 The Emergence and Development of Language


LEARNING OUTCOMES

By the end of this section, you will be able to do the following:


• Describe the communicative abilities of wild animals such as birds and primates.
• Distinguish primate communication from human language.
• Identify the biological features of early hominins that were central to the emergence of language.
• Identify the archaeological evidence for the emergence of language.

There are some seven thousand languages spoken in the world today. Most people are proficient in at least one
of them, possibly more. But people are biologically capable of mastering any of them, and have been since
birth. Humans are born language ready. For a human baby, any language will do. With passive exposure to
language (simply hearing it without any formal instruction), human toddlers learn the complex rules and vast
vocabularies of the language spoken (or signed) around them. This astounding feat is made possible by
specific biological features in the brains and bodies of human babies, features designed to help them
understand and produce language. The learning of language then triggers further changes in our brains,
making possible certain kinds of reasoning and thought as well as communication with others.

Access for free at openstax.org


6.1 • The Emergence and Development of Language 181

FIGURE 6.2 When teaching language to their children, some parents teach signs (such as those of American Sign
Language) as well as spoken words for objects. The theory is that sign language and spoken language are processed
in different parts of the brain. Teaching these two forms of language together may provide deeper cognitive
reinforcement and greater chance of recall. This baby is making the sign for “bird.” (credit: “Bri signs ‘Bird’” by Bev
Sykes/flickr, CC BY 2.0)

Drawing on biological and archaeological evidence, researchers seek to understand how, why, and when
humans developed the biological features associated with language and, once language emerged, how the
practice of language changed the way of life of early humans. Language became a building block for human
culture of increasing complexity. Innovations such as stone tools, hunting, and using fire for heat and cooking
were made possible by language. In turn, these new skills enhanced the survival of those who practiced them,
increasing the likelihood that those people would live to pass on their genetic makeup to their offspring. This
means that certain biological features were key to the invention of human culture and that human culture was
key to the biological development of humans. We think of this as a reciprocal system of biocultural coevolution.
Put another way, biology and culture developed in tandem, with language as the link between the two.

No one really knows when or how humans invented language. The problem is that language, whether spoken
or gestural, leaves no direct trace in the archaeological record. Lacking direct evidence, researchers must be
creative, combining various indirect forms of evidence to suggest theories about how language may have
begun in humans. Based on such methods, researchers think that language may have emerged between
50,000 and 200,000 years ago. The largeness of this window of possibility is due to the indirect nature of the
evidence and a great deal of controversy about which elements may have been most important in the process
of language development. In this section, we look at these forms of indirect evidence, starting with
communication in the animal kingdom.

Animal Communication
All animals communicate with each other and even with other species (Tallerman and Gibson 2011). Many use
vocalizations like calls, growls, howls, and songs. Many also use gestures such as dances, postures, and facial
expressions. Some change the color of their scales, skin, or fur. Some produce strong-smelling body fluids
sprayed in their environment or rubbed on their own bodies. All of these activities are used to tell other
animals about territory, food sources, predators, and mating opportunities.
182 6 • Language and Communication

FIGURE 6.3 Canada geese fly in a V formation to conserve energy and to keep track of all the birds in the
formation. Coordination and communication are essential for the group. (credit: “Canada Geese” by Alex Galt, US
Fish and Wildlife Service/flickr, CC BY 2.0)

Many people might be tempted to think that animals speak to each other just as we do, that their various forms
of communication are roughly equivalent to language. Does your dog bark and jump excitedly whenever you
pick up the leash? Isn’t that a way of saying, “C’mon! Let’s go for a walk!”

Some forms of animal communication are fairly simple, such as this canine leash mania. Others are far more
complex, such as the way an octopus can change the color of and patterns on its skin for hunting, courtship,
and camouflage. Fireflies use bioluminescence to attract mates and as a defense mechanism. Some fish
generate electric fields to advertise their species and sex. Many animals use a vast lexicon of postures and
gestures to communicate messages to one another and even to other species. When a bird issues a predator-
alert call, squirrels respond as well. Many mammals pay attention to the predator warnings of birds.

Are these complex forms of communication equivalent to language? Take a closer look at one famous example
of complex animal communication and compare it to human language.

A Waggle is Not a Word: The Complexity of Language


Consider the famous “waggle dance” of the honeybee. Upon finding a good source of nectar such as a grove of
wildflowers, a worker bee returns to the hive and performs a special flight pattern consisting of a figure-eight
waggle followed by a return loop alternating right and left. The direction and duration of the waggle
communicate the direction and distance to the location of the desirable food source (Seeley 2010; Frisch
1993).

Access for free at openstax.org


6.1 • The Emergence and Development of Language 183

FIGURE 6.4 Diagram of the waggle dance of the honeybee. The movements performed by the bee during this
dance communicate the direction of and distance to a food source to its fellow hive members. (credit:
“20180622-FS-WashingtonDC-KTC-024” by Kelly Chang, US Forest Service/flickr, Public Domain)

The waggle dance is certainly a complex and effective form of communication, but does it qualify as language?
Communication refers to the transfer of information from a sender to a receiver. Communication can be
voluntary or involuntary, simple or complex. Language is a specific, complex, systematized form of
communication involving the use of vocal or gestural units (words or signs) that can be combined and
recombined in larger structures (sentences) that can convey an infinite array of complex meanings. Language
is a form of communication. Not all communication is language.

Central to the infinite possibilities of language is a set of rules that govern just how sounds, signs, words, and
phrases may be combined. These rules structure the order of words, dictating, for example, where to put
subjects and actions in an utterance so that listeners will be able to find them. Rules also tell us whether words
indicate a single thing or multiple things and whether actions occur in the past, present, or future. Complex
forms of animal communication such as the waggle dance do contain some systematic rules governing the
sequence, duration, and intensity of certain segments of the communication, but they are highly constrained
to very limited contexts. For example, the waggle dance can be used to signal nectar sources near and far, but it
cannot be used to discuss the weather or comment on the laziness of the queen. Unlike the relatively “closed”
systems of communication common among animals, human language is open-ended. Our languages have the
distinctive quality of allowing actors to combine units in an infinite number of ways to produce new meanings.

Simple Signs and Pant-Hoots: Language in Primates


Biological anthropologists posit that we share a common ancestor with the other great apes (gorillas,
chimpanzees, bonobos, and orangutans) about five to eight million years ago. As nonhuman primates do not
produce language in the wild, the biological and cultural features that promoted language must have emerged
after that. However, studies aimed at teaching human language to nonhuman primates have revealed that
individuals of these species are able to master basic vocabulary and use simple words and word combinations
to obtain the things they want. So the great apes must have some biological features that enable them to learn
human language in a partial and limited way.

You may have heard of Koko, the gorilla famous for learning to use sign language. Sign language is used in such
studies because nonhuman primates lack the distinctive vocal tract required to make the sounds of human
language. Researcher Penny Patterson taught Koko to use about a thousand signs, roughly the vocabulary of a
three-year-old child (Patterson and Linden 1981). Patterson reported that Koko could comment on things that
were not currently present in her environment, such as personal memories. According to Patterson, Koko
could joke and lie and teach other gorillas to sign. She could even invent new signs. Many of these claims are
disputed by other researchers. Some point out that the evidence is largely anecdotal and relies on the
interpretation of Patterson herself, hardly an objective observer. Though controversial, Patterson’s path-
184 6 • Language and Communication

breaking work with Koko provided a wealth of data and opened up new possibilities for understanding the
language abilities of nonhuman primates.

FIGURE 6.5 Koko learning to play the guitar. Koko became famous for learning to communicate with humans using
roughly 1,000 signs taught to her by researcher Penny Patterson. (credit: “ODCnewBegin9” by FolsomNatural/flickr,
CC BY 2.0)

Human-reared chimps, gorillas, bonobos, and orangutans have all been taught to use gestures or tokens to
refer to things in the world around them, often combining those signs in a rule-based way to make comments
and requests. Even though many linguists are skeptical of these studies, the use of symbolic systems in
cooperative interactions to achieve goals does seem to indicate that great apes have the basic capacity to
generate some sort of protolanguage. Protolanguage refers to a very simple set of gestures or utterances that
may have preceded the development of human language. But do apes display these abilities due to some innate
capacity or because we have taught them symbolic systems? Perhaps learning a symbolic system has changed
the brains of these individual animals in distinctive ways.

FIGURE 6.6 Chimpanzees use gestures and facial expressions as well as vocalizations to communicate with one
another. (credit: “Chimpanzees” by foshie/flickr, CC BY 2.0)

Many primatologists conduct research on the vocal and gestural forms of communication used by primates in

Access for free at openstax.org


6.1 • The Emergence and Development of Language 185

the wild, looking for those biological features that might underpin the human capacity for language. Wild
chimpanzees, for instance, produce a wide range of calls, including hoots, pant-hoots, pant-grunts, pant-barks,
rough-grunts, nest-grunts, alarm barks, waa-barks, wraas, screams, and soft panting play sounds (Acoustical
Society of America 2018). Primatologists have listened closely to these calls. Some argue that chimp
vocalizations are not much like human language, as calls are fairly fixed and limited in their meanings.
Chimps may use a rough grunt to indicate a food source, but they do not seem to have specific grunts for
specific food types. Monogamous pairs of gibbons, a smaller species of ape, are known to perform elaborate
morning duets. Gibbons have an array of predator calls as well. Research comparing duets with predator calls
suggests that gibbons compose their songs to convey specific information, each note carrying a certain
meaning (Clark et al. 2006). While impressive, the ability to manipulate notes to convey a limited range of
meanings is still a far cry from the infinite productivity of human language. The limitless recombination of
signs that produces the flexible, open-ended quality of language is missing in the communication systems of
wild primates.

Human Biology and the Emergence of Language


There must be something special about us to make possible the distinctively flexible and open-ended
communication system of language. Research has focused on our throats, our brains, and our genes, looking
for the biological features that allowed for the emergence of language.

The Vocal Tract


Humans have evolved a very unusual vocal tract with a descended larynx (otherwise known as the “voice box”)
and a large and rounded tongue positioned in the mouth to enable a remarkable array of sounds (Lim and
Snyder 2015). Some researchers suggest that our throats may have evolved in response to walking upright or
changes in diet or a combination of those two factors. Humans also have more deliberate control over
breathing than nonhuman primates. In order to better understand when hominins developed this distinct
vocal apparatus, researchers examine the hyoid bones of hominins to see if they resemble those of modern
humans. The hyoid is a U-shaped bone in the human throat that helps us swallow and move our tongues. The
few hyoids that have been found in the fossil record suggest that our distinctive vocal tract may have been
developed around 500,000 years ago. This means that Neanderthals likely had the same vocal abilities as
modern humans.

FIGURE 6.7 Evolutionary changes in the vocal tract enabled the development of spoken language in humans. The
image on the left shows the vocal structures an early ancestor to humans. The image on the right shows the vocal
tract of modern humans. The position of the vocal structures in the early ancestor allows for eating and breathing at
the same time. The position of these structures in modern humans allows more sounds to be produced and more
words to be spoken in sequence. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

Brain Structure
Several features of the human brain are considered prerequisites to language, including the overall (large) size,
the division into specialized hemispheres, and certain structures like Broca’s and Wernicke’s areas. Broca’s
area is a region of the brain associated with the production of speech. Wernicke’s area is essential to the
comprehension of language. Both are most often located in the left hemisphere of the human brain (for left-
186 6 • Language and Communication

handed people, both can be located on the right side). How did we acquire these brain features so essential to
language? A great deal of controversy surrounds this question, as researchers debate when and how these
structures evolved.

FIGURE 6.8 The locations of Broca’s area and Wernicke’s area in the human brain. Broca’s area, responsible for
the articulation of speech, is next to the motor area, where the movements of the body are controlled. Wernicke’s
area, associated with language comprehension, is situated beside the primary auditory area, where sounds are
processed. (credit: “1605 Brocas and Wernickes Areas-02” by OpenStax College/Wikimedia Commons, CC BY 3.0)

Most recently, research has focused on “mirror neurons,” special brain cells that seem to enable mimicry
(Lim and Snyder 2015). Many researchers think that the ability to understand the actions of others and
recreate those actions ourselves is a fundamental prerequisite for language. That is, in order to be able to talk
to each other, early hominins must have been able to evaluate and interpret each other’s actions and
reproduce them in similar contexts. In primates like monkeys, scientists have discovered a system of
specialized neurons called the “mirror neuron system” that enables primates to recognize and imitate actions.
Monkeys and apes cannot talk, but they can recognize, interpret, and imitate actions performed by other
primates. The neurological studies that revealed mirror neurons are too invasive to perform on humans, but
neuroimaging studies suggest that a similar mirror neuron system does exist in humans.

FIGURE 6.9 Mirror neurons are most likely involved in the spread of contagious yawning. Mirror yawning happens
between humans and can even happen across species. You can make your dog yawn! (credit: “Sleepy” by
Toshimasa Ishibashi/flickr, CC BY 2.0)

Brain imaging studies on humans have located evidence for the mirror neuron system in a region of the brain

Access for free at openstax.org


6.1 • The Emergence and Development of Language 187

close to Broca’s area. So it is possible that the mirror neuron system inherited from primates provided a
foundation for the later emergence of a brain structure devoted to language production in hominins. If
imitation and language are in fact connected in this way, then a system of gestures may have paved the way for
the development of language. Some researchers now hypothesize exactly this: that hominin language evolved
from a system of gestures to a system of vocalizations.

The “Language Gene”


In the late 1980s, medical researchers became aware of a particular speech disorder common among
members of one family in West London. Many members of this family could not pronounce words. Many
stuttered. Many had very limited vocabularies. Geneticists traced the disorder to a genetic mutation on
chromosome number 7 of the human genome. (See Biological Evolution and Early Human Evidence for more
on chromosomes and genes.) The mutation was located on a gene named FOXP2, prompting some researchers
to dub this “the language gene.” Some hypothesize that FOXP2 may have played a role in the development of
language in humans (Lim and Snyder 2015).

At first, researchers thought that only humans had the FOXP2 gene, but subsequently a form of this same gene
has been identified in many vertebrates, including mice, bats, fish, and songbirds. In mice, the gene appears to
be related to vocalizations. In birds, it seems to be linked to birdsong. All primates have FOXP2, but the human
copy is slightly different than that of nonhuman primates. Some researchers think this mutation occurred
around 260,000 years ago and may have enabled the development of spoken language in Neanderthals and
Homo sapiens.

Other researchers are skeptical of the notion that one gene could be responsible for the emergence of spoken
language (Tallerman and Gibson 2011). Many anatomical developments and cognitive processes—connected
to different parts of the human genome—are involved in human language. These developments and changes
would have required mutations in other parts of the genome of early Homo. While the mutation of FOXP2 in
Homo may have played a role in language development, other mutations would have been important as well.

Hominin Material Culture


Evidence from the material culture of hominins such as Homo habilis and Homo erectus is also used to
speculate about the emergence of human language. Early hominins developed stone tool technologies and
created stunning works of art. The production and use of such tools and artwork must have required a
complex set of social and cognitive abilities. Those same types of social and cognitive skills are important to
human language. It is possible that language emerged as part of a whole complex of material culture.

Archaeological evidence and linguistic theory come together in a model suggesting that the invention of tools
by early hominins was linked to the invention of language. Some linguistic theorists suggest that the
evolutionary changes in brain structure that allowed for the development of tool use also support the
emergence of language. Furthermore, the innovations of tools and language are entwined in a reciprocal
relationship; evolutionary pressure to develop tools stimulated the development of language, and the
development of language facilitated increasingly complex tool making and tool use.

There are two theories to explain the connections between advances in tool use and language. The first rests
on the assumption that tool making requires a considerable degree of cognitive planning. You cannot make a
useful tool by just picking up a rock and randomly chipping away at it. Hominins like Homo habilis and Homo
erectus must have known just what kind of rocks would work as base and chipper and how to execute a set of
precise chips in a certain sequence to achieve a sharp blade without breaking the core. The mental processes
important to this sort of planning are hypothesized to have also enabled hominins to do the sort of quick
planning involved in the production of complex speech (Tallerman and Gibson 2011).

A second theory linking tool use and language emphasizes the importance of imitation in passing along the
complex set of skills involved in tool making. Neuroscientist Michael Arbib suggests that the ability to imitate
may have generated the first gestural language among hominins (2011). And he has developed a model to
describe how imitation and tool making may have evolved together over time. About 2.5 million years ago,
Homo habilis began making basic stone choppers, cores with flakes removed, used for butchering carcasses.
Such choppers are called Oldowan tools, named after the site in Olduvai Gorge in Tanzania where they were
188 6 • Language and Communication

first found. Arbib has theorized that the production of Oldowan tools required the ability for hominins to
imitate each other’s actions. Simple imitation would make it possible for a learner to reproduce the actions of
an accomplished tool maker through observation and mimicry. This ability to imitate is biologically rooted in
the system of mirror neurons discussed earlier. As hominin brains acquired the ability of simple imitation
involved in tool production, they might also become capable of the kind of gestural communication we see in
apes today—not language, but a precursor to it. Investigate this diagram for more about the evolution of
language (https://openstax.org/r/researchgate).

The array of action-oriented mirror neurons, tool innovation, and language all progressed together in hominin
evolution. As tool technology developed, Homo erectus began making distinctive pear-shaped hand axes about
1.6 million years ago. A more intricate form of imitation would have been necessary to teach this sort of tool
making to others, corresponding to the emergence of protolanguage. This protolanguage might have been a set
of simple one-word utterances corresponding to concepts such as “yes,” “no,” “here,” or “there.”

We don’t have any hominin brains to examine, but remember that in the human brain, the system of mirror
neurons is assumed to be situated near Broca’s area, which is associated with human speech. So very likely,
protolanguage emerged in the same part of the brain as the ability to imitate. The explosion of innovations in
tool making over the past 100,000 years is linked to the emergence of complex human language. While the
development of mirror neurons and the ability to learn tool making required biological changes to the brain,
Arbib argues that the last step, the emergence of language, was purely cultural.

6.2 Language and the Mind


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Describe the role of language in categorizing items in the natural world.
• Explain the Sapir-Whorf hypothesis.
• Provide at least two examples of linguistic universals.
• Describe how metaphor shapes how we talk about abstract concepts.

As discussed in the previous section, certain cognitive abilities were crucial to the development of language in
humans. And reciprocally, once language emerged, it shaped our thoughts and actions in ways that helped our
species cooperate, invent, learn, and adapt to the environment. Language must have been a fundamental
element in the creation of human culture (singular) and the eventual development into human cultures
(plural) as different groups of humans moved into different geographical areas and began adapting to different
conditions.

One key advantage of language is that it provides a way of encoding specific information about the
environment and sharing that information with others so that it endures over time. If, say, there are snakes in
an area, it would certainly be important to distinguish the venomous ones from the harmless ones, so probably
there would be separate words for those two categories of snake or at least words for each specific snake so
that people could alert each other to the presence of a dangerous one.

This means that early language must have been developed relative to environmental conditions. Linguistic
anthropologists are interested in the way that language varies across cultures, reflecting different
environmental, historical, and sociocultural conditions. This is called linguistic relativity.

On the other hand, languages are also constrained by human anatomy and cognitive abilities. Say there were
two species of snake in an area, one poisonous and the other harmless, but you could not tell them apart by
looking at them. (This is actually an adaptive strategy deployed by harmless animals called adaptive mimicry.)
In this case, early humans probably would have had just one word for snake, indicating that sometimes a
snake’s bite made you sick and sometimes it didn’t. As this example shows, the human visual apparatus
shapes our understanding of the world, which, in turn, shapes our language.

Consider another example from the natural world—the beetle. There are over 300,000 types of beetles in the
world. How many can you name? All of them? Ten of them? Two of them? Outside of written scientific
taxonomy, there is no language in the world that contains separate terms for each kind of beetle. This is not

Access for free at openstax.org


6.2 • Language and the Mind 189

only because there are only a few thousand of each type of beetle living in any one environment but also
because of limits to the number of terms any person can learn and remember. Our vocabulary is constrained
by the limits of human memory.

FIGURE 6.10 So many beetles. How many can you name? There are over 300,000 types of beetles in the world.
Outside of written scientific taxonomy, no language in the world contains separate terms for each kind of beetle.
(credit: “display drawer 3” by Joana Cristovao, Natural History Museum/flickr, CC BY 2.0)

So language is shaped not only by environmental conditions but also by how humans interact with their
environments. Our common human anatomy influences our comprehension of the world, and that
comprehension is expressed in language. This insight suggests that all languages must have some things in
common by virtue of the fact that all humans have the same anatomy and cognitive abilities. Some linguistic
anthropologists are interested in discovering these linguistic universals.

In the next section, we take a look at some intriguing research on both linguistic relativity and linguistic
universals, seeking to better understand how language interacts with our human minds.

Linguistic Relativism and the Sapir-Whorf Hypothesis


As seen in previous chapters, it was lamentably common for scholars in the early 20th century to think of non-
Western societies as backward and primitive, incapable of complex, abstract thought. Franz Boas worked hard
to disprove these racist notions, seeking to demonstrate the equal sophistication of all peoples and cultures.
Boas trained a student named Edward Sapir who was particularly interested in how non-Western languages
conveyed forms of complex, abstract thought that were different from the Euro-American habits of thought.
Sapir, in turn, trained a student named Benjamin Whorf who further elaborated on this theme in his own
research (Ahearn 2017). The result is what we have come to call the “Sapir-Whorf” hypothesis.

The Sapir-Whorf hypothesis argues that the particular language you speak influences how you think about
reality (Lucy 2001). Thus, different languages encourage different habits of thought. This is an essential tenet
of linguistic relativity. Whorf based his argument on a comparison between the Native American language of
Hopi and what he called “Standard Average European” (SAE), a broad category of European languages
including English. Whorf was interested in how speakers of each language might think differently about time.
In English vocabulary, time is divided into units that can be counted. English speakers talk about the number
of seconds, minutes, or days before an event or consider the number of months or years since something
190 6 • Language and Communication

occurred. In Hopi, according to Whorf, time is conceived as indivisible and enduring, a whole process
unfolding. The Hopi talk about the flow of events in a completely different way, a processual way Whorf termed
“eventing.” Whorf argued that these linguistic features profoundly influenced sociocultural life in each of these
two contexts. Holding with the understanding of time as process, Hopi culture emphasized preparation,
endurance, and intensity. Coordinating with the SAE expression of time as countable units, Euro-American
culture emphasized schedules, accounting, and record keeping. Many people use a calendar to keep track of
meetings, appointments, and assignments. Whorf would argue that the English language encourages us to
think of time and events in this way, as a spatialized set of boxes to be filled up with discrete objects.

In connection with the Sapir-Whorf hypothesis, it is sometimes said that the “Eskimos” have 400 words for
snow. This notion is both problematic and untrue. The first problem is that “Eskimo” is considered a
derogatory term by the Inuit and Aleutian peoples to whom it has been applied. And, secondly, the claim turns
out to be wrong. Anthropologist Laura Martin (1986, also described in Ahearn 2017) has debunked the myth by
documenting that Arctic peoples really have just two root words for talking about snow, one for snow that is
falling and the other for snow that is on the ground. They use these roots much as English-speakers would, to
talk about snowstorms, snowflakes, snow drifts, and snow melt. The Sapir-Whorf hypothesis is not typically
applied to the vocabularies of different cultures anymore.

Recall the earlier example about snakes. We hypothesized that a culture might not distinguish between two
species of snakes if those snakes looked identical. But if people gradually came to notice that the poisonous
snakes were always found in trees while the harmless snakes were always found on the ground, it is likely that
a different term would come to be used for the tree-dwelling kind of snake, the one with the harmful bite. That
is to say, even if a culture previously had only one term for snake, the people in that culture could easily
understand that there were, in fact, two kinds and would be able to change their language to mark that
difference in their vocabulary for future reference. Their vocabulary would not limit their thinking to such a
degree that they could not conceive of two different kinds of snake.

Rather than specific vocabulary words, researchers who study linguistic relativity have come to focus on larger
abstract topics like space. In languages such as English, when people want to tell someone where a particular
object is, they most frequently use language focused on their own bodies. English-speakers say, “You have a bit
of arugula on the left side of your mouth” or “Grab the pink top hat on the shelf above you.” This way of talking
relies on the human body as a point of reference and therefore is relative to the bodies of the speaker and/or
hearer. This creates confusion when the speaker is facing the person they are talking to, sometimes prompting
someone to say, “No, my left, not your left!” Steven Levinson has conducted research on languages that do not
use the human body to talk about direction at all (2003). Instead, they use the cardinal directions (north, south,
east, west) and specific features of their environments (mountains, oceans) to talk about where things are. A
speaker of the Australian indigenous language of Guugu Yimithirr might say, “Watch out for the snake just
north of your foot!” This way of talking about space is absolute, not relative. Such speakers never have to say
“No, my north, not your north,” as there’s only one absolute north. Research suggests that these different ways
of reckoning give us different kinds of mental maps, such that a Guugu Yimithirr speaker might be better at
absolute navigation than an English speaker, and perhaps more adept at finding her way back home if she lost
her way.

Linguistic Universals and Folk Taxonomies


While linguistic relativists explore how different linguistic patterns shape different thought patterns (and vice
versa), other linguists are interested in how all languages are constrained by our common human biology and
in finding universal linguistic patterns. There are specific domains of language that lend themselves
particularly well to this kind of inquiry. One of them is color. The reason for this is that color relies directly on
our human visual system, invariant across cultures.

And yet there is enormous diversity in the ways that different cultures divide up the spectrum of possible color.
Some cultures have hundreds of color terms, while others have only two or three. Researchers Brent Berlin and
Paul Kay analyzed the color term systems of 98 languages and found that the diversity of color term systems is
governed by one set of rules. All of these color term systems are comprised of a few basic colors with specific
colors added to the scheme over time (Kay 2015, Berlin and Kay 1969). The color schemes of all cultures are

Access for free at openstax.org


6.2 • Language and the Mind 191

based on the distinction between black and white (or light and dark). If a culture has only two terms, those two
terms will always be black and white. The next most important color is red. If a culture has three color terms,
those terms will be black, white, and red. Next comes green and yellow, then blue, then brown, then purple,
pink, orange, and gray, always in that order. Berlin and Kay suggested that these rules form a pattern for the
way all languages develop over long periods of time. Although the scheme proposed by Berlin and Kay has
been revised a little in the past 50 years, the basic tenets have held up pretty well (Haynie and Bowern 2016).

FIGURE 6.11 Berlin and Kay’s developmental scheme for the elaboration of color terms. Some cultures only
distinguish black from white. When another term emerges, that color is red. After that, green and yellow are added,
either one first. Then blue and brown are added, in that order, and then one of these four: purple, pink, orange, or
gray. (attribution: Copyright Rice University, OpenStax, under CC BY 4.0 license)

VIDEO
Vox: The Surprising Pattern behind Color Names around the World
Click to view content (https://openstax.org/r/surprisingpattern)

Oddly, though this finding lends very strong support to the notion of linguistic universals, the very same
research has also been used to argue for linguistic relativity. Paul Kay later teamed up with another linguist,
Willet Kempton, to consider how different color schemes might affect how people “see” color in the
environment around them (1984). They presented people with color chips on the spectrum between true blue
and true green. They asked subjects how they would group all the colors into two categories. People who spoke
languages that had terms for both blue and green drew a more distinct boundary between the two colors than
people who had just one word for both blue and green.

Clearly, relativity and universalism are both aspects of human language. Our common biology plays a role in
how humans interact with the world, providing regularity to the way all languages categorize not only color but
also plants, animals, weather, and other natural phenomena. Researchers who study the systems of categories
people use to organize their knowledge of the world have a term for those cultural systems: folk taxonomies.
The folk taxonomy for any area of human knowledge reflects both human biology and the surrounding
environment and sociocultural practices. There are folk taxonomies for plants, animals, clouds, foods, and the
cries of babies.

Folk taxonomies are not just vocabulary terms; they frequently structure any kind of distinction that is
meaningful within a culture, even those that rely on simple qualifiers like “good” and “bad.” One example is
death, surely invariant across cultures. Societies all over the world distinguish between a “good” death and a
“bad” death. These notions reflect cultural beliefs and values—such as the American notion that a good death
is a painless one. Among the Akan peoples of Ghana, a good death is the death of someone who has led a very
long life, achieving all of the culturally valued accomplishments in life, such as getting married, having
children, accumulating property, and providing support to friends and family members (Adinkra 2020).
Imagine a very old great-grandmother surrounded by her many descendants as she lies in her bed, heaving
one final breath as she drifts away peacefully into death. That is a good death. A bad death is tragic and violent,
the sudden death of a person who has not had the chance to really live a full life. Think of a young person
drowning or dying in a traffic accident. That is a very bad death. If someone has had a good death, that person
is eligible to become an ancestor if the correct rituals are performed. The body must be washed, publicly
mourned, and buried in a beautiful casket in a public cemetery, often with grave goods like tools and money to
help the person in the afterlife. Ancestors are important, as they watch over their living relatives, possibly
192 6 • Language and Communication

helping them out if called upon through libation or other ritual means. If someone has had a bad death,
however, they may become an angry ghost, haunting family members with bad luck. The funeral rites of bad
deaths are rushed, minimal, and private in order to avoid commemorating or communicating with the agitated
spirit.

Categorization is central to our perceptions, thoughts, actions, and speech. The way humans categorize objects
and experiences is limited by the way our brains and bodies work, resulting in linguistic universals like the
developmental scheme of color terms. However, the complex meanings associated with cultural categories
vary widely, resulting in a great deal of linguistic relativity. Linguistic relativism and universalism are often
described as opposite positions, but in fact, they are both essential and complementary features of human
language.

Meaning and Metaphor


How are you feeling today? Are you feeling up or feeling down? If you’re feeling low, try doing something fun to
lift your spirits. Take care of yourself so you don’t fall into a depression.

An old theory suggested that languages are primarily referential; that is, each language contains a set of
vocabulary terms that correspond to elements in the natural world. According to this theory, language
functions as a mirror of reality. We have seen in the last section, however, that different languages divide up
the natural world in different ways, from the natural domains of color and plants to the human domains of life
and death. Moreover, humans use language to talk about abstract issues like mood, social relationships, and
communication itself. It is fairly easy to use our terms for spatial organization to talk about the location of
concrete objects like arugula on somebody’s face. But what about more abstract issues? How do we talk about
becoming friends with someone? How do we discuss an argument we’re making in a term paper? How do we
talk about how we’re feeling today?

Mood is like color insofar as the human physiology of mood structures a set of near-universal basic categories
including happiness, sadness, anger, fear, disgust, and surprise. And yet, because mood occurs on a spectrum,
it is divided up in different ways by different cultures. Consider “schadenfreude,” a German word combining
the roots for “damage” and “joy.” Schadenfreude refers to taking pleasure in another’s misfortune. There is no
equivalent word in English.

We don’t just use language to identify the emotions we’re feeling. We also talk about the process of developing
an emotion, how one mood leads to another, and how we can prevent ourselves from feeling a certain way.
These are mysterious and abstract processes. How do we do this? We use metaphor. A metaphor is a linguistic
idiom where use what we know about something concrete to think and talk about something abstract.
Cognitive linguists George Lakoff and Mark Johnson argue that metaphor is the primary way we create
complex meaning in language (1980). In terms of mood, we use our concrete language of direction to talk
about our abstract experience of mood. A positive mood is understood as up, while a negative mood is
considered down. If you’re feeling really happy, you might say you’re on top of the world. If you’re really sad,
you might say you’re down in the dumps. In fact, the word for prolonged sadness, depression, literally refers to
a sunken place or the act of lowering something.

Metaphor is one of those things that you don’t notice until you start paying attention to it. And then you realize
that it’s everywhere: in the way you think about time, number, life, love, physical fitness, work, leisure, sleep,
and thought itself, just to name a few highly metaphorical topics. Just about any abstract area of experience is
structured by metaphorical thinking. Here are three common metaphors in English, with examples.

LIFE IS A JOURNEY

He took the wrong path in life.


As you move ahead, you should follow your dreams.
When I left home, I came to a crossroads in life.
If you work hard, you’ll arrive at a sense of accomplishment later in life.

LOVE IS SWEET

She’s my sweetheart.

Access for free at openstax.org


6.3 • Language, Community, and Culture 193

The newlyweds went on a honeymoon.


Sugar, would you pass the salt?
Our love was sweet, but then it went sour.

ARGUMENT IS COMBAT

The candidate launched a personal attack against her opponent.


His position on taxes is indefensible.
Armed with facts, she won the argument.
His criticism really hit the mark.

There are thousands and thousands of metaphors in English. Many abstract domains rely on a combination of
various metaphors used to describe different aspects of the experience. You can think of love as sweet (as
above) but also as a journey (as in “Will the couple go forward together, or will they go their separate ways?”) or
as combat (as in “He slew me with his come-hither glance”).

Metaphor is found in all human languages. Some specific metaphors, like the directional metaphors used to
describe mood, are found in many, many cultures. A study by Esther Afreh (2018) found that the king of Asante
(in Ghana) frequently uses metaphorical language in his public speeches, including such familiar ones as “life
is a journey,” “life is a battle,” “ideas are food,” “knowing is seeing,” and “death is sleep.” Though the speeches
were delivered in English, Afreh notes that these metaphors also exist in Akan, the local language of the Asante
people. Alongside her analysis of the English-language speeches, she notes many proverbs and phrases in
Akan that use the same metaphors.

As with our discussion of categorization in the last section, metaphor is both relative and universal. Lakoff and
Johnson argue that our common human biology structures our experiences of things like emotion and life.
When you’re feeling really sad, you might literally feel like lying down, and when you’re really happy, you might
jump with joy. We may use the notion of a journey to structure our understanding of life, social relationships,
and time in general because in our everyday life, we move forward in space to pursue objects and activities.

Sometimes the reasons for cross-cultural similarities are not so directly linked to human biology. English and
Chinese have similar metaphorical systems for talking about moral issues. In both languages, the adjective
meaning “high” is associated with things that are lofty, noble, or good, while the adjective “low” is used to
describe things that are mean, contemptible, or evil (Yu 2016). Alternatively, it is also possible in both
languages to describe moral behavior as “straight,” while immoral behavior can be termed “crooked.”

On the other hand (to deploy a useful metaphor), different cultures do rely on different metaphors to talk about
some domains of experience, metaphors that emphasize certain aspects of those abstract topics. Consider the
English notion that “time is money.” This is a metaphor, pure and simple, but many English speakers believe it
to be absolutely true. You can spend time, waste time, save time, and invest time. So time does seem like
money in capitalist cultures. But time is not literally money. Nor is time a journey or a horizontal line in space,
though these are common ways of thinking about time in the English language. Time is just time, an abstract
idea. Certainly Whorf did not find the Hopi talking about time as money. English speakers think of time in
terms of money because they live in a society in which time is treated as money, a society that tends to
monetize nearly everything, from land and labor to advice, attention, and even body parts like human sperm.

6.3 Language, Community, and Culture


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Explain the role of culture in the acquisition of language.
• Describe how language can form the foundation of sociocultural groups in speech communities.
• Describe how people code-switch among speech communities.

While language is critical to individual human thought, its basic function is to communicate messages in
human communities. That is, language is fundamentally social. Through social interaction, humans learn the
language of their community. And through language, humans express community identity and coordinate
194 6 • Language and Communication

their activities.

Language Acquisition and Language Socialization


Imagine that someone handed you a babbling baby and said to you, “Teach this baby the basic rules and values
of our culture.” What would you do?

Likely, you’d start by teaching the baby your language. Without language, it’s pretty hard to teach rules and
values (unless you are a really good mime). Luckily, babies come into the world with special cognitive abilities
that make them ready to learn language. Most babies undergo a rapid process of language learning between
the ages of nine months and three years. Babies proceed through a set of stages that allow them to learn
language just by being exposed to surrounding talk. Many scholars study the problem of language acquisition,
examining precisely how humans manage to learn language in a diversity of sociocultural contexts.

So your babbling baby would probably learn language just by being exposed to it. But what if someone wanted
to hasten the process or make sure their baby was particularly excellent with language?

An American would probably interact with the baby in a particular way, sitting the baby on their lap facing
them, pointing to objects and asking basic questions in a quiz-like fashion. “See the cookie? Where did the
cookie go? In my tummy!” The person might say these types of things while talking in a high-pitched, sing-
song voice. Linguists call this type of talk “motherese.” In many other cultures, caregivers do not interact with
babies in this way. In some cultures, oversimplified “baby talk” is considered detrimental to language learning.
The context of language learning might involve a whole host of characters beyond the baby and the caregiver,
encompassing all household relatives, neighbors, visitors, and even strangers. Language is not always “taught”
to babies, but is often witnessed and overheard. Rather than quizzing her baby American style, a mother in
Kaluli society in Papua New Guinea is more likely to sit her baby on her lap facing outward, talking “for” the
baby in conversations with siblings (Ochs and Schieffelin [1984] 2001). In West Africa, babies spend large parts
of the day wrapped on the backs of their mothers where face-to-face interaction with her is impossible. But
they overhear the talk around them all day long, and people frequently engage their attention in brief
interactions. In the field of language socialization, researchers go beyond the various stages of language
learning to focus on the social contexts in which language is acquired. As social contexts shape the way
children learn language, language itself becomes a means of learning about sociocultural life.

Whether facing their caregivers or facing out to the social world around them, babies in all cultures learn to be
proficient in their languages. And yet, in American culture, the notion persists that language proficiency relies
on very precise forms of interaction between caregiver and baby, the American model of motherese. Every
culture has specific ideas about language, how it is acquired, how it varies across social groups, how it changes
over time, etc. These ideas are termed language ideologies. Some of these ideas, like the notion that babies
have a special “window” of opportunity for learning language, are supported by linguistic research. Others,
however, are challenged by ethnographic and cross-cultural research.

Speech Communities and Code Switching


A ten-year-old girl described one of her stuffed animals as “derpy.” Here is a snippet of her conversation with
her mother:

Thisbe: Look at his face. He’s so derpy.


Jennifer: Derpy? I don’t know that word. What does it mean?
Thisbe: Like, kind of stupid. Kind of dumb.
Jennifer: Oh, ok. Like Clover [our dog], when she fell off the couch. Was that derpy?
Thisbe: No, that’s not derpy! It’s like ... Mom, I just can’t explain it to you. You just have to know.

All speakers of a particular language form a hypothetical community, sharing a common grammar and
vocabulary, as well as a set of understandings about how language is used in different situations. Within this
large group are smaller groups of speakers who use the common language in special ways unique to that
group. Anthropologists use the term speech community to describe such a group (Muehlmann 2014). Speech
communities often have distinctive vocabularies, grammatical forms, and intonation patterns. Using these
features appropriately, members of the speech community demonstrate their membership in the group.

Access for free at openstax.org


6.3 • Language, Community, and Culture 195

The concept of speech community was originally used to describe the distribution of dialects in a language. A
dialect is a form of language specific to a particular region. For instance, in the Philadelphia metropolitan
area, it’s common for local people to pronounce the word “water” as “woohder,” as if it nearly rhymes with the
word “order.” It’s also common to use the phrase “yooz” for the second-person plural (as in, “Yooz better drink
some woohder!”). Linguists William Labov, Sharon Ash, and Charles Boberg famously mapped out these
dialectical differences in different regions of the United States (2006). Over time, a dialect can accumulate
such unique linguistic features that it develops into a separate language. Indeed, the distinction between a
well-developed dialect and a language is largely political. Nation-states may downplay regional differences as
mere dialects in order to maintain linguistic unity, while separatist political movements may champion their
way of speaking as an entirely different language in order to justify their demands for independence.

FIGURE 6.12 Map of American dialects. While English is the official language in all areas of the United States, the
particular way it is spoken varies from region to region. (attribution: Copyright Rice University, OpenStax, under CC
BY 4.0 license)

Other researchers have focused on the speech communities of ethnic groups and immigrants. Researchers use
the term vernacular to describe dialects that are not necessarily regional but associated with specific social
categories, such as groups based on ethnicity, age, or gender. Anthropological research on African American
Vernacular English (AAE), Chicano English, and Native American English have all shown how these
vernaculars shape distinctive forms of storytelling, arguing, and criticism (Chun and Lo 2015). Rather than
seeing ethnic vernaculars as “incorrect” forms of English, researchers demonstrate how vernaculars like AAE
are highly structured linguistic systems with regular grammatical patterns and innovative vocabularies (Labov
1972a). In formal settings like American classrooms and courtrooms, these alternative ways of using English
are too often stigmatized as lazy, unintelligent, or just plain wrong. Believing their own English to be the
“correct” form, authority figures often forbid the use of alternative vernaculars of English and refuse to engage
in any effort to understand those forms.

More recent research on vernaculars has explored how speakers maneuver among the styles of language they
encounter in their daily lives, engaging in various languages, dialects, vernaculars, and other elements of style.
We all use a variety of linguistic styles, and many speak more than one language. Addressing different
audiences, U.S. President Barack Obama used linguistic strategies to “Whiten,” “Blacken,” “Americanize,” and
“Christianize” his public identity, thus subverting racial stereotypes and indicating his membership in a
diversity of communities (Alim and Smitherman 2012). In parts of the world that were previously colonized by
Europeans, European languages have been maintained as the formal language of government and education
even as most people speak local languages in their everyday interactions with kin, neighbors, merchants, and
other community members. In these postcolonial contexts, people tack back and forth between various styles
196 6 • Language and Communication

of their local languages as well as shifting between the local language and the European one. Such strategic
maneuvering among linguistic styles, called code-switching, is done by people in many difference contexts.

For many people, the style of language spoken in elite settings such as schools and government institutions
has the effect of disempowering and marginalizing them. Linguistic anthropologists examine how vernaculars
associated with elite and professional groups become a means of in-group solidarity and out-group exclusion.
Anthropologist and lawyer Elizabeth Mertz (2007) conducted participant observation in first-year classes at
several American law schools, looking at how law students are taught to “think like a lawyer.” Using a version of
the Socratic method, law professors teach their students to set aside the moral and emotional elements of
cases to view them purely as texts subject to abstract, professional analysis. The ability to master the linguistic
maneuvering and arcane vocabulary of this form of analysis becomes a prerequisite for becoming a lawyer.
The American justice system is thus dominated by people who are trained to set aside humanistic concerns in
favor of textual authority and manipulation. Mertz’s study shows how people are socialized by language
throughout their lives, not just in childhood. And it alerts us to the way that language can be used to elevate the
learned perspectives of elites, dismissing the moral and emotional perspectives of others.

6.4 Performativity and Ritual


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Provide examples of the performative function of speech acts.
• Describe how ritual language can be performative.
• Identify the informal ways that people “talk back” to formal speech.

The Performativity of Language: Speaking as Action


Consider the following pairs of sentences. What are the differences between the two sentences in each case?

1. Boris and Natasha are married.


2. Boris and Natasha, I now pronounce you husband and wife.

1. Natasha: Boris lost his job.


2. Natasha: Boris, you’re fired!

1. Boris: Natasha, I ate the last pickle.


2. Boris: Natasha, I apologize for eating the last pickle.

In all the above pairs, the first sentence is a report about an event. The second sentence makes an event
happen. In the sentences about the pickle, the second sentence does not make the pickle disappear, but it does
create an apology for that action, hopefully altering the consequences of the pickle-eating. In the previous
section, we explored how we use language to think and reason about the world around us. This is an essential
function of language, but it is not the only one. We also use language to do things—that is, to perform actions in
the world.

Way back in the 1930s, Bronislaw Malinowski explored how people use language in culturally specific ways to
play an active part in their societies (Duranti 2012). Malinowski described how the people of the Trobriand
Islands used magical language to compel the growth of yams, bananas, taro, and palms in their carefully
cultivated gardens. Magical spells, like all ritual language, aim at making something happen through the
special manipulation of public speech. We see the same use of language in other ritual settings like marriages
and naming ceremonies. The plot of many a Hollywood romantic comedy hinges on the moment the partners
say “I do” and the officiant pronounces them married. In American marriage ceremonies, it is clear that ritual
language is the tool that marries people—not the rings, or the pageantry, or the blessings of family and friends,
or any other aspect of the ritual.

In his influential book How to Do Things with Words (1962), philosopher of language J.L. Austin coined a term
for action-oriented language: performatives. The most obvious performatives use phrases like “I pronounce,”
“I order,” “I promise,” “I warn,” or “I appoint.” Sentences that begin with these phrases are explicitly uttered
with the intention of doing something through the act of speaking. As he dug deeper into the performative

Access for free at openstax.org


6.4 • Performativity and Ritual 197

function of language, however, Austin realized that performatives are not so much a separate category of
utterances but an aspect of most of the things we say. Even when people are making a simple descriptive
statement, they are saying it for a reason. The power of speech to make things happen is called
performativity. Consider the following sentences:

The exam is next week.


The dog is pawing on the door.

The above sentences are statements about an event or situation. However, if a professor announces to the
class, “The exam is next week,” this is not merely an observation, but a warning—a cue to students to study in
preparation for the upcoming exam. And if someone tells their roommate, “The dog is pawing on the door,”
they are essentially telling that person to let the dog out.

Like metaphor, performativity is one of those aspects of language that permeates everyday speech. Once you
learn about it, you recognize performativity in just about everything you say. Spend a few hours paying
attention to each utterance as you go about your activities. You’ll find that you rarely use language to merely
describe what’s going on. You speak in order to generate a response or result, even when you just say “Hi.”

The Performativity of Ritual Language


Just as Malinowski studied the special language used in garden magic among the Trobrianders, many
contemporary linguistic anthropologists study the role of performative language in various ritual settings. In a
recent article, Patience Epps and Danilo Paiva Ramos examine the performance of incantations among the
Indigenous Hup community of the northwest Amazon (Epps and Ramos 2020). An incantation is a patterned
set of phrases or sentences used to compel a magical result. Among the Hup, incantations are used by elders
for protection, healing, and causing harm. While Epps and Ramos were conducting fieldwork in the area, Hup
elders expressed concerns that the young men in the village were not learning the repertoire of important
incantations properly, thus endangering the health and safety of the community. The elders invited Epps and
Ramos to write down their incantations for healing and protection in order to preserve them for future
generations. Epps and Ramos documented and analyzed these incantations in consultation with Hup elders.

In the article, Epps and Ramos analyze an incantation used by the Hup to protect travelers on paths through
the rain forest. This incantation is recited by an elder before a group of Hup people embark on a journey. After
providing the original text and its English translation, Epps and Ramos describe the incantation’s structure
and poetic features, including the use of metaphor and repetition of phrases. As a whole, the incantation lists
various dangers and helpful entities and enacts certain magical practices through the speech itself. At the
beginning of the incantation, the elder states that he is enclosing the entire path in a protective “canoe,” much
as a traveler on a river would ride in a canoe. This canoe is named after a particular snake, the mussurana
snake (Clelia clelia), a constrictor snake that eats other snakes and is immune to their venom. Thus, the
incantation is creating a metaphorical shield of protection around the travelers, making them safe from
venomous snakebites. In the second section of the incantation, the elder lists all classes and subtypes of
snakes that might be encountered in the rain forest, asserting a kind of taxonomic mastery over the snakes.
Summoning the snakes one by one, he tells of lining them up, sitting them down, and feeding them sticky coca
and tobacco. The snakes then sit quietly, their jaws stuck together by the sticky substance so that they are
unable to bite anyone. The incantation goes on to deal with several other malevolent entities and engage with
beneficial entities to help the travelers in their journey.

Informal Back-Talk: Teasing, Grumbling, and Gossip


Linguistic anthropologists most frequently rely on long periods of fieldwork, living in the communities they
study and witnessing and even participating in ritual events where performative language is deployed. Such
events include protection and healing magic, but also naming ceremonies, puberty rites, weddings, funerals,
and other rituals that mark the passage of persons from one social status to another. Anthropologists term
such rituals “rites of passage” (discussed in detail in Anthropology of Food). At such ritual events, elders or
religious specialists are called upon to perform the ritual language necessary to publicly move persons from
the previous category to the new one.

Naming ceremonies are a great example of the power of performative language in rites of passage. In many
198 6 • Language and Communication

West African societies, a baby is not considered a true person until they have been publicly named by an elder
or religious official in a naming ceremony performed a certain number of days after the baby is born. Extended
family and friends attend the ceremony as markers of their relationship to the baby. Guests bring gifts such as
rice and cloth for the baby, and they are rewarded for their attendance with prepared food and kola nuts.

During his fieldwork in southeastern Senegal, linguistic anthropologist Nicholas Sweet witnessed the naming
ceremony for a baby in a Pular-speaking village (2019). When the family were gathered in the compound of the
baby’s father, the imam rose, faced east, gave the blessings of the prophet, and then performed the naming of
the baby girl (in Arabic, English translation below):

In the name of God, the gracious and the compassionate


O Allah, send blessings on our master Muhammad
O Allah, send blessings on our master Muhammad
O Allah, send blessings on our master Muhammad
The name of the child has come here, her mother and her father have named her Aissatou
The name of the child has come here, her mother and her father have named her Aissatou
This is what was written on the tablet of Allah
May God grant her blessings

While carefully recording the formal performative language so important to this naming ceremony, Sweet was
also attuned to the more informal kinds of language that surrounded the main action. For instance, just before
the imam’s performance, some friends of the family were gathered around the baby, remarking on her beauty.
As a way of showing their admiration, some of the men joked and teased one another about the prospect of
marrying her someday. Other relatives teased the baby’s parents with demands for kola nuts and other food.
As dramatically performative as the official naming was, this informal language was also performative,
providing a way for guests to socially configure their various relationships to the new person in their
community.

Someone important had been left out of the ceremony—the great-aunt of the baby, also named Aissatou. As the
baby was her namesake, Auntie Aissatou had been invited and should have been a featured guest at the
ceremony. But when the time came to perform the ceremony, she had not arrived yet, and so they went on
without her. Afterward, as guests were making their way home, they crossed paths with Auntie Aissatou, who
was just then on her way to the event. Realizing that the naming had already been performed, she complained
that she had been waiting for someone to fetch her and bring her to the ceremony. Auntie Aissatou was angry
that she had missed the ceremony as well as the gifts distributed afterward.

Wrapping a scarf around her head in imitation of an imam, Auntie Aissatou continued on to the compound of
the baby’s father. Striding ceremoniously into the compound, she addressed a number of elders still gathered
there. In a parody of the official naming performance, she faced east, delivered the blessings of the prophet,
and then announced:

The name of the child has come here. It is Buubu Nooge (Trash Owl).

The audience of relatives erupted in laughter but also protest, interrupting Auntie Aissatou to correct her with
the baby’s true given name. But Auntie Aissatou persisted, saying over and over again that the baby’s name had
come and it was “Trash Owl.”

Why Trash Owl? In this community, it is believed that witches turn themselves into owls when they fly through
the night. “Trash” seemed to refer to the joke gifts of garbage (broken flip-flops, an old sock) in a small gourd
that Auntie Aissatou presented in lieu of the usual baby gifts of food, cloth, and soap.

In the days following the naming ceremony, the teasing name for the baby became a running joke in the
community, especially among people who had not been invited to the ceremony but felt that they should have
been. In order to quash the teasing nickname, the baby’s family was compelled to make a number of visits
around the community with appeasing gifts of kola in an effort to get everyone to recognize the baby’s proper
name. Once Auntie Aissatou and the others had received their visits and kola, they abandoned the name Trash
Owl, recognizing the baby as Aissatou, the namesake of Auntie Aissatou.

Access for free at openstax.org


6.5 • Language and Power 199

This incident illustrates the power of parody and gossip to steal performative power from the authoritative
realm of formal speech, giving excluded and marginalized people a way to “talk back” to authority. There are
many ways of doing this. Often, audiences to formal speech will deliberately misunderstand or creatively
interpret the proclamations of authority figures.

6.5 Language and Power


LEARNING OUTCOMES

By the end of this section, you will be able to:


• Explain how language can operate as a gendered form of power.
• Identify how racial categories and bias are expressed through linguistic practices.
• Describe strategies used by communities to revive their dormant languages.

Gender and Language


In 2018, the word “mansplaining” was added to the Merriam-Webster dictionary. The word is defined as “what
occurs when a man talks condescendingly to someone (especially a woman) about something he has
incomplete knowledge of, with the mistaken assumption that he knows more about it than the person he's
talking to does” (“Words We’re Watching” 2018).

The word was inspired by an article written in 2008 by the feminist blogger Rebecca Solnit. In the article “Men
Explain Things to Me,” Solnit described an incident at a party in which she mentioned to a man that she had
recently written a book about a particular photographer. Immediately, the man interrupted to inform her about
a very important book that just came out about that same photographer, a book he had read about in The New
York Times. After the man had described the book in great detail, Solnit’s friend finally intervened to say that
the book he was talking about was, in fact, written by Solnit. In the wake of Solnit’s article, other women writers
described similar experiences in their workplaces, schools, and relationships, and the whole phenomenon
came to be called “mansplaining.”

FIGURE 6.13 Rebecca Solnit, the author of the article “Men Explain Things to Me”. The term “mansplaining” has
become a well discussed topic in the years since she introduced the term. (credit: “Rebecca Solnit” by Charles
Kremenak/Wikimedia Commons, CC BY 2.5)

Have you ever witnessed mansplaining? Have you ever mansplained to someone? Embedded in the very term
is a notion about gender and language. The idea is that men and women have different styles of speech, styles
that reflect and reinforce inequality between genders.

In recent years, many writers have pushed back against the term “mansplaining,” arguing that all men do not
always speak this way to all women. Some argue that many men are much more respectful and sensitive to the
200 6 • Language and Communication

dynamics of power in their conversations with women. Some argue that privileged White women tend to speak
in a mansplaining way to male waiters and salespersons or to people of color more generally. Others suggest
that older people speak in a condescending way to younger people, or vice versa.

Have you ever become annoyed with a friend or relative who repeatedly interrupts you? Have you ever noticed
how some people tend to end their sentences with rising intonation, making everything they say sound like a
question? How about a person who ignores what you say but then rephrases your idea and takes credit for it?
Many people associate these ways of speaking with gender, the way men speak or the way women speak. As
noted in the discussion of language acquisition, every culture has ideas about how language operates, called
language ideologies. The idea that American men and American women have distinctive styles of speech is a
language ideology. Whether it is true or not is a question for linguistic research, but this idea has become a
widespread way of thinking about gender, power, and language in American culture.

In the 1970s, linguists inspired by the women’s movement turned their attention to the way gender shapes
different patterns of speech. In her influential book Language and Woman’s Place (1975), Robin Lakoff argues
that women and men are socialized to speak in distinctive ways that empower men and subordinate women.
Lakoff describes women’s speech as uncertain, excessively polite, and full of hedges, emotional language,
euphemism, and tag questions (“Don’t you think?”). Other linguistic researchers have found that men tend to
interrupt women far more than vice versa, even when the women speaking are doctors and the men are their
patients (Zimmerman and West 1975, West 1998).

Building on this research, Deborah Tannen generalized beyond speech patterns to describe two entirely
different communicative subcultures for American men and American women (1990). When men and women
speak to one another, Tannen argues, they are speaking cross-culturally, deploying different motivations and
expectations for talk. Men engage in conversation to assert their status in a social hierarchy, while women are
more interested in building solidarity through social connection. Men authoritatively report information to
their interlocutors, while women engage in conversational rapport with their interlocutors. In popular media,
differences in the speech styles of men and women are frequently linked to purported differences in specific
parts of male and female brains, such the corpus callosum, the amygdala, and the hippocampus. In this way,
gendered patterns of speaking are naturalized as biological.

Like the pushback against the term “mansplaining,” researchers have begun to challenge the view that women
and men are embedded in different linguistic subcultures with different patterns of speech, motivation, and
interpretation. Psychologist Janet Hyde conducted a meta-analysis of hundreds of quantitative studies to see if
widespread notions about gender and language were actually borne out by linguistic data (2005). Along with
notions of power, Hyde was interested in testing the idea that women are chattier and more deferential than
men. Focusing on studies of children, Hyde found that boys and girls exhibited no differences at all in reading
comprehension, verbal reasoning, and vocabulary. The tendency for boys to interrupt or speak assertively was
only very slightly higher than for girls. The girls’ tendency toward self-disclosure and cooperation with their
conversation partners was only slightly higher than for boys. The only significant differences Hyde found were
in smiling and correct spelling (girls did more of both).

How do we reconcile research demonstrating differences in the way men and women talk with data that
suggests very little difference in the speech patterns of girls and boys? One could argue that children have not
been entirely socialized into their assigned gender category. Perhaps the discrepancy suggests that gendered
ways of speaking are cultural, not biological, and that, for children, the most intense period of socialization is
yet to come in adolescence.

Moreover, ethnographic research by linguistic anthropologists shows that patterns of speech associated with
men and women are culturally relative. Reversing the American stereotypes, anthropologists working in
Madagascar and New Guinea have found that women are expected to speak in a more confrontational and
argumentative style, while men are associated with more cooperative, euphemistic, and ceremonial speech
(Keenan [Ochs] 1974, Kulick 1992, both cited in Ahearn 2017).

So both quantitative and ethnographic research overturn the notion that women and men are biologically
engineered to use language in different ways. That leaves us with the conclusion that any differences in the
ways men and women talk are entirely cultural. Literary scholar Judith Butler argues that gender identities are

Access for free at openstax.org


6.5 • Language and Power 201

not biological but are performed through language and other cultural practices, particularly those centered on
the body (1988). So when men and women speak in certain ways, they are socially performing their gender
identities, whether consciously or unconsciously. Moreover, through their linguistic performances, people
enact their own versions of gender in complicated ways that transcend the neat dichotomy of male and female.
You probably have a language ideology that tells you how men and women speak in your culture, but do you
always speak in the style associated with your assigned gender category? Nobody does. And some people
rarely do. As these contradictory performances build up over time, the very notion of gender can change.

PROFILES IN ANTHROPOLOGY

Kira Hall
1962-

Area of Anthropology: Kira Hall’s (https://openstax.org/r/colorado) work is situated at the intersection of


sociolinguistics and linguistic anthropology. In graduate school, she studied with Robin Lakoff in the
linguistics department at the University of California–Berkeley, earning her PhD there in 1995. For her
dissertation, she examined the linguistic strategies of Hindi-speaking hijras in Banaras, India. Hijras are
members of a third-gender group in many Indian communities. Most hijras were raised as boys and later
adopted the intersex behaviors and language of the hijra identity. Hall analyzed how hijras navigated aspects
of gender embedded in Hindi, such as certain verbs and adjectives that are marked as feminine or masculine.
She showed how hijras alternate between these gendered forms, code-switching as a reflection of their own
ambiguous identities. She explored how hijras use obscene forms of language to shame people into giving
them money. She showed how they had developed their own secret language as a way of communicating with
one another, signaling their identity to others, and excluding non-hijras from understanding their
conversations.

Accomplishments in the Field: Reflecting her work at the boundaries of linguistics and anthropology, Hall has
held academic positions in the anthropology department at Yale University and the linguistics department of
Stanford University. Currently, she is professor of linguistics at the University of Colorado at Boulder, with a
joint appointment in the anthropology department. She is also director of the Program in Culture, Language,
and Social Practice at UC-Boulder. Since 2019, she has served as the president of the Society for Linguistic
Anthropology of the American Anthropological Association.

Importance of Their Work: Hall’s work highlights how language operates within hierarchies of gender,
sexuality, and socioeconomic class. In addition to her work on hijras, she has published articles on language
and sociality in autism, female mass hysteria in upstate New York, and Donald Trump’s use of gesture and
derisive humor in the 2016 Republican Party primaries.

Race and Ethnicity


On many government forms, people are asked to identify their “race.” Forms in the United States often include
five categories: Black, White, Asian, American Indian/Alaska Native, and Native Hawaiian/Other Pacific
Islander. The category “Hispanic or Latino” is often listed as an ethnicity rather than a race. On the 2020 U.S.
Census, people were presented with 14 racial categories to choose from: White, Black or African American,
American Indian or Alaska Native, Chinese, Filipino, Asian Indian, Vietnamese, Korean, Japanese, Other Asian,
Native Hawai’ian, Samoan, Chamorro, and Other Pacific Islander. Again, “Hispanic, Latino, or Spanish” was
listed as a question of “origin.” Even with so many options, many Americans still could not find a category that
represented their racial or ethnic identity.

As you’ll remember from earlier chapters in this text, race is not biological. There is no accurate way to divide
up the gradual spectrum of human biological variation, meaning that biological categories of race are entirely
imaginary. However, we also know that social categories of race are very powerful tools of discrimination,
subordination, solidarity, and affirmative action. Earlier in this chapter, we studied how sets of categories, “folk
taxonomies,” are embedded in language. We saw how different cultures divide up the natural world differently.
202 6 • Language and Communication

Likewise, race and ethnicity are folk taxonomies, embedded in language and organizing the social world into a
neat set of groups. These categories are real insofar as they have shaped the structure of our society,
advantaging some groups and disadvantaging others. And they are real insofar as they shape our thoughts and
actions and even our subconscious habits and tendencies.

Like gender, race and ethnicity are performed in language. We use language in conscious and unconscious
ways to express racial and ethnic belonging as well as exclusion. Take the use of Spanish catchphrases by
Americans who do not speak Spanish. Many Americans intend to be jokey and fun by using Spanish phrases
such as “hasta la vista!” and “no problemo” as well as deliberately incorrect ones such as “buenos nachos” and
“hasta la bye bye!” Anthropologist Jane Hill found that middle-class, college-educated White Americans were
most likely (among other Americans) to use this “mock Spanish” (2008). People who use these phrases
consider them harmless and even respectful, while Spanish speakers are often insulted by the association of
Spanish with silliness. Hill argues that such phrases are only funny because they covertly draw from
stereotypes of Spanish speakers as foolish, lazy, and inept.

Similar arguments about cultural appropriation and stereotyping can be made about the use of Black
vernacular speech by White Americans. In the United States, a variety of English called African American
English (AAE), or African American Vernacular English, is spoken by many people in predominantly Black
communities. With the widespread popularity of Black culture, many White Americans have picked up phrases
and grammatical features of AAE while knowing very little about the vernacular and the people who speak it as
their primary language. To many Americans, AAE is just imperfect English (it is not, as we’ll see in a moment).
So what are White people signaling when they say things like “chillin’,” “lit,” “on fleek,” “aa’ight” (for “alright”),
“ima” (for “I’m going to”) and “Yasss, Queen!” Does the use of this language convey respect for the communities
associated with Black vernacular English? Or does it demean and subordinate Black Americans who speak
AAE?

People who use mock Spanish and mock AAE typically do not mean to insult anyone. The problem is not one of
intent, but of context. In American culture, most middle-class White people speak forms of English considered
standard or mainstream (Lippi-Green 2012). In fact, Standard American English (SAE) is historically based on
the language of Anglo American immigrants. The adoption of White Anglo-English has always been considered
critical to successful assimilation by minority and immigrant groups. Success at complete assimilation is often
measured by the ability to speak SAE without an accent. But SAE is not speaking “without an accent.” SAE is an
accent—the accent of White people whose ancestors emigrated from the British Isles.

SAE is the dominant language of American public spaces, including schools, workplaces, government, and
media. People who speak SAE without effort or accent can speak freely in these spaces, knowing that their
language will be understood and respected. Americans whose primary language is Spanish or AAE often
struggle to be understood and taken seriously in American public life. Given this context, it can seem
disrespectful for White Americans to appropriate Spanish and AAE as tools of humor while denigrating and
marginalizing the actual speakers of these languages.

The issue is further complicated by the widespread and persistent notion among White Americans (and many
Black Americans too) that AAE is not a language at all, but merely a hodgepodge of slang and bad grammar.
This view is simply wrong, another language ideology that has no basis in fact. AAE is a rule-governed form of
English with its own regular system of sounds, grammar, and vocabulary (Labov 1972b). For historical reasons,
AAE shares many features with the English spoken by White southerners in the United States as well as
working-class Cockney English from London (Ahearn 2017). Rooted in historical experiences of slavery and
segregation, Black Americans have developed their own distinctive set of innovative linguistic features to
supplement the more basic structure of American English. Consider the following three sentences:

He is angry.
He angry.
He be angry.

The first sentence is SAE, and the second and third are AAE alternatives. In SAE, this conjugation of the verb
“to be” describes a situation happening in the present. But the SAE present tense of “to be” is a bit vague, as it
can mean “right now, this very minute” or a more ongoing situation, perhaps describing a person who is

Access for free at openstax.org


6.5 • Language and Power 203

frequently or enduringly angry. AAE helpfully distinguishes between these two possibilities. “He angry” means
angry “right now,” whereas “He be angry” indicates a more ongoing situation. In linguistic terminology, the
second example is called “copula deletion” and the third is called “the habitual be.” Both are used in regular
ways to indicate the difference between momentary and enduring conditions.

AAE is governed by many more rules and features that provide its speakers with expressive possibilities not
available to speakers of SAE. In other words, AAE is not only a rule-bound vernacular; it’s a more developed
and complex form of English. Linguists have been trying to convey this message to the American public since
the 1970s (Labov 1972a). Read more about AAE at the Anti-Racism Daily website (https://openstax.org/r/Anti-
Racism).

Rather than recognizing the innovative contributions of vernaculars like AAE, language policy in the United
States stigmatizes non-SAE vernaculars as “bad English” spoken by uneducated and unintelligent people.
Linguist John Baugh calls this “linguistic profiling” (2003). With colleagues Thomas Purnell and William
Idsardi, Baugh (1999) compared the response of California landlords to apartment inquiries spoken in SAE,
AAE, and Chicano-American English (CAE). In Woodside, California, landlords responded to SAE inquiries 70.1
percent of the time. Inquiries in AAE received responses only 21.8 percent of the time and CAE inquiries only
28.7 percent of the time. Research in American schools and courtrooms corroborates the discriminatory
effects of linguistic profiling on access to housing, education, and justice.

The use of language to discriminate and marginalize is certainly not limited to American English. Elites in
many cultures define their own way of speaking as “correct” and “official,” using linguistic practices in public
spaces to disempower other groups based on class, race, ethnicity, gender, and sexuality. How can people
respond to these forms of linguistic marginalization? For many upwardly mobile speakers of “nonstandard”
vernaculars and languages, the process of becoming successful has involved the abandonment of their
primary way of speaking in favor of standard, elite forms of language privileged in public discourses. But there
is another alternative. As speakers of nonstandard vernaculars and languages move into public discourses,
they can hold on to their primary languages, code-switching from context to context. Some language activists
celebrate the genius of their “home” languages and work to nurture and revive them, as we will see in the next
section.

Can speakers of dominant languages contribute to the process of celebrating and revitalizing marginalized
languages? Is it always insulting or racist for speakers of a dominant language to use phrases from another
vernacular or language? Some people think so. Certainly it is harmful to use phrases that reference negative
stereotypes (even indirectly). But what if your limited use of a few phrases can help you communicate with
someone from a different background? What if SAE speakers started quoting Spanish or AAE in ways that
highlight positive aspects of those speech communities? What if White people started learning AAE in order to
publicize the genius and complexity of this American vernacular? What if you learn another language or
vernacular in order to subvert the forces of cultural segregation in your own society? There are no easy
answers to such questions.

Endangered Languages: Repression and Revival


In 1993, a Wampanoag (https://openstax.org/r/tribalpedia) woman living on a reservation in Cape Cod,
Massachusetts, had a mysterious dream, recurring on three consecutive nights (Feldman 2001). In the dream,
a circle of Wampanoag were singing in a language she did not understand. When she woke, words of the
language stuck with her, and she longed to find out what they meant. Were these words of Wôpanâak, the
language of her ancestors? Wôpanâak had died out in the mid-1800s.

The woman was Jessie Little Doe Baird, a social worker and mother of five. Haunted by those words, she began
reading through documents from the 1600s written in Wôpanâak, including letters, deeds to property, and the
earliest translation of the Bible printed in the Western hemisphere (Sukiennik 2001). Though frustrated in her
efforts to find the meaning of her dream words, she developed a passion for the language of her ancestors and
began working with local Wampanoag communities to reclaim their common language of Wôpanâak.
Community response was enthusiastic. Committed to the project, Baird went to MIT to study linguistics,
earning a master’s degree. Based on her survey of Wôpanâak documents, she wrote a dictionary and began
teaching Wampanoag students to speak the language (https://openstax.org/r/Women_of_the_Century).
204 6 • Language and Communication

By learning their ancestral language, Baird and her students found themselves reconnecting with Wampanoag
culture in unexpected ways. The grammar of Wôpanâak, for instance, puts the speaker at the end of the
sentence rather than the beginning. Whereas English speakers would say “I see you,” Wôpanâak speakers
would say something like “You are seen by me.” Baird suggests that this word order highlights the value of the
community over the individual, putting awareness of the other ahead of the self. Wôpanâak displays
alternative logic in the formulation of nouns as well. For instance, in English, animal names reveal little or
nothing at all about the animal. The words “cat,” “mouse,” and “ant” are based on arbitrary sounds that convey
no information about their referents. In Wôpanâak, however, animal names frequently contain syllables that
refer to the animal’s size, movement, and behavior. The word for “ant,” for instance, incorporates syllables
communicating that the animal moves about, does not walk on two legs, and puts things away.

By now, you know that forms of cognition and culture are embedded in language. The languages of the world
encode diverse experiences of time, space, life, death, color, emotions, and more. A language serves as a form
of oral documentation of the surrounding environment, a survey of the flora, fauna, topography, and climate of
an area. Forms of cultural wisdom are preserved in the stories and proverbs of a language. History is recorded
in epic tales and legends. Language can be essential to maintaining cultural identity, affirming the common
history and values of a people while providing them with a distinctive way of communicating with one another.

Among the seven thousand languages spoken in the world today, roughly 40 percent of them are in danger of
dying out in the next hundred years. A language is considered dead when it is no longer spoken by any living
person. Wôpanâak was once considered a dead language. Some linguists argue that no language should ever
really be considered “dead,” however, and prefer the terms “dormant” or “sleeping.” So long as there are
written or audio records of a language, it can come to life again, a process called language revitalization.
Returning to a language that has become dormant or endangered, community members can develop strategic
programs to spread, nurture, and modernize the language, ensuring it has a future for generations to come.

Languages generally become endangered or dormant through processes of colonialism and imperialism. In
North America, as Native Americans were forcibly removed from their lands and confined to reservations in
the 1800s, they were compelled to send their children to boarding schools where they were forbidden to speak
their Native languages or practice their Native cultures. As foreign settlers seized lands in Australia, New
Zealand, and Hawaii, they established similar schools, aimed at assimilating Indigenous children by stamping
out their language and culture. Elsewhere, more gradual processes of endangerment can occur when a new
language offers opportunities for employment and trade only available to speakers of that language. Parents
may encourage their children to learn the new language in order to take advantage of these opportunities, and
children may come to reject their own language as a backward language of old people.

Many, many languages have risen from dead or comatose states, among them Cornish, Hawaiian, Hebrew,
Scots-Gaelic, the Ainu language of Japan, the Indigenous Australian language of Barngarla, the Indigenous
New Zealand language of the Māori people, and the Native American languages of the Navaho and Blackfoot
peoples. Often, as with Wôpanâak, the impetus for language revival comes from energetic community
members who feel the loss of their language as a threat to their cultural survival. These concerned people
create programs to document the language and teach it to children and adults. They establish contexts where
the language is spoken routinely and exclusively. Sometimes they work with linguists to develop these
programs.

The most successful of these revitalization strategies are immersion schools and master-apprentice programs.
In the early 1980s, Māori language activists developed full-immersion preschools, called Te Kōhanga Reo, or
“language nests” (King 2018). In these nests, very young children are taught language and culture by Māori
elders—grandmothers and grandfathers in the community. Native Hawaiians have developed a similar
program of language nests, called Pūnana Leo. Early on, some parents worried that children in immersion
schools would not learn the dominant national language well enough to be successful in later life, but research
has shown that such children do just as well or better in later classroom performance and standardized
testing. Many language revitalization projects combine early immersion with later bilingual education (Hinton
2011, 2018). The Navaho Immersion School in Arizona provides immersion education for the first three years
of schooling and then introduces English as the medium of instruction through grade seven. From grades eight
to twelve, children are taught in Navaho half the time and English the other half.

Access for free at openstax.org


6.5 • Language and Power 205

FIGURE 6.14 Sign in front of a full-immersion school in Seatoun, New Zealand. All classes are held in the Māori
language. (credit: “Te Kura Kaupapa Maori O Nga Mokopuna” by Tom Law/flickr, CC BY 2.0)

One of the challenges of school-based revitalization programs is finding enough adults sufficiently proficient
in the language to teach it to children. Among the strategies of language revitalization that target adult learners
is the master-apprentice approach. The original Master-Apprentice Language Learning Program was founded
in California by the Advocates for Indigenous California Language Survival (Hinton 2018). The strategy has
since spread all over the world. In these programs, a proficient speaker and a motivated learner spend 20
hours a week together, using the target language plus gestures and other nonverbal communication to engage
in various activities.

When successful, language revitalization can empower individuals and energize communities. Learning their
heritage language, people come to understand the distinctive genius and complexity of their culture while
preserving a crucial means of transmitting that culture across generations.

MINI-FIELDWORK ACTIVITY

Dispute Analysis

Choose a friend, relative, or acquaintance with whom you might disagree on a particular issue. Suggested
issues might include musical taste, what makes a good restaurant, how to behave on a date, the best form of
physical exercise, or anything else you feel comfortable talking about but might disagree on. Ask the person if
they would consent to being recorded for an anonymous fieldwork exercise. If so, record a 5-to-10-minute
conversation with that person in which you discuss the issue. Then, review the conversation. What seem to be
the goals of the two interlocutors? What is the pattern of turn taking? What truth or knowledge claims are
made by each speaker, and what are the bases of those claims? How is authority constructed and challenged?
How does each one respond to the assertions of the other? How does the conversation turn out in the end?

Suggested Readings
Ahearn, Laura. 2017. Living Language: An Introduction to Linguistic Anthropology. 2nd ed. Chichester, West
Sussex, UK; Malden, MA: Wiley-Blackwell.

Duranti, Alessandro. 1997. Linguistic Anthropology. Cambridge, UK: Cambridge University Press.
206 6 • Key Terms

Key Terms
code-switching the practice of tacking back and which language is learned as well as the role of
forth between various linguistic styles depending language in social learning.
on contexts and interlocutors. linguistic relativity the way that language varies
communication the transfer of information from a across cultures, reflecting different
sender to a receiver; can be voluntary or environmental, historical, and sociocultural
involuntary, simple or complex. conditions.
dialect a form of language specific to a particular linguistic universals common elements found in
region. all human languages, attributable to human
folk taxonomies systems of categories that people anatomy, perception, and cognition.
use to organize their knowledge of the world. metaphor a linguistic idiom using something
FOXP2 a gene on chromosome number seven that concrete to think and talk about something more
is found in many vertebrates; sometimes called abstract.
“the language gene” because the human mutation mirror neurons special brain cells that seem to
seems to be associated with language. enable mimicry.
incantation a patterned set of phrases or naming ceremony a public ritual that officially
sentences used to compel a magical result. grants personhood by bestowing a name.
language a complex, systematized form of performativity the functional power of language to
communication involving the use of vocal or make things happen.
gestural units (words or signs) that can be protolanguage a very simple set of gestures or
combined and recombined in larger structures utterances that may have preceded the
(sentences) that can convey an infinite array of development of human language.
complex meanings. Sapir-Whorf hypothesis the theory that the
language acquisition the process of learning a particular language you speak influences how you
language. think about reality.
language ideologies specific ideas about language speech community a community of speakers
that are widespread in a culture, including how sharing a common grammar and vocabulary, as
language is acquired, how it varies across social well as a set of understandings about how
groups, how it changes over time, etc. language is used in different situations.
language revitalization the process of reviving an vernacular dialects that are not necessarily
endangered or dormant language using strategies regional but associated with specific social
such as immersion schools and master- categories such as groups based on ethnicity, age,
apprentice programs. or gender.
language socialization the social contexts in

Summary
Language and culture are closely entwined in the particularly relative, influencing how children learn
evolutionary development and contemporary languages in various sociocultural contexts as well
diversity of human societies. Human language as how people use language to create speech
differs from animal communication in its communities. In ritual contexts, language is used
complexity and flexibility, aspects of human performatively to accomplish social action as well as
communication made possible by unique human challenging those actions. As a tool of power,
biological and genetic features. The complexity of language structures gender, race, and ethnic
language makes it a powerful tool in shaping human dynamics. Recognizing the fundamental importance
thought, providing categories and metaphors for of language to the preservation of culture, many
organizing our information about the world. Though Indigenous communities have developed strategies
language shapes thought and action in universal to revive their heritage languages using immersion
ways, many aspects of language vary widely relative schools and master-apprentice programs.
to local cultures. The social aspects of language are

Critical Thinking Questions


1. What might humanity be like if humans had never developed language? What social and

Access for free at openstax.org


6 • Bibliography 207

cultural forms would not be possible without 3. List the speech communities to which you
language? How would we survive? Would we be belong. Do all members of a speech community
capable of creating tools or art? Would our social share exactly the same vocabulary and practices?
relationships be different? Would our social Do speech communities overlap?
groups be different? 4. Aside from weddings, list rituals in your culture
2. Describe a romantic relationship, one you have that rely on the performance of language. How do
experienced or observed. How did it begin, people use forms of commentary and back talk to
develop, endure, or end? Now, make note of how reshape those performances?
many times in your description you relied on 5. How does language operate as a form of power in
various metaphors. Is it possible to fully describe schools and universities? Consider the gendered
romance without the use of metaphor? Do these norms of language as well as racial and ethnic
metaphors shape the way we think about dynamics.
romance?

Bibliography
Acoustical Society of America. 2018. “Can Chimpanzee Vocalizations Reveal the Origins of Human Language?
While Closely Related to Humans, Researchers Discover that Chimpanzees' Vocalizations Resemble Human
Language Less than You'd Expect.” ScienceDaily. May 8, 2018. www.sciencedaily.com/releases/2018/05/
180508081505.htm.

Afreh, Esther Serwaah. 2018. “Metaphors Otumfo? Lives By: A Cognitive Linguistic Study of Metaphors in
Some Addresses by Otumfo? Osei Tutu II, Asantehene.” Cognitive Semantics 4 (1): 76–103.

Ahearn, Laura. 2017. Living Language: An Introduction to Linguistic Anthropology. 2nd ed. Chichester, West
Sussex, UK; Malden, MA: Wiley-Blackwell.

Alim, H. Samy, and Geneva Smitherman. 2012. Articulate While Black: Barack Obama, Language, and Race in
the U.S. New York: Oxford University Press.

Arbib, Michael. 2011. “From Mirror Neurons to Complex Imitation in the Evolution of Language and Tool Use.”
Annual Review of Anthropology 40:257–73.

Austin, J. L. 1962. How to Do Things with Words: The William James Lectures Delivered at Harvard University
in 1955. London: Oxford University Press.

Baugh, John. 2003. “Linguistic Profiling.” In Black Linguistics: Language, Society and Politics in Africa and the
Americas, edited by S. Makoni, G. Smitherman, A. F. Ball, and A. K. Spears, 155–68. New York: Routledge.

Berlin, Brent, and Paul Kay. 1969. Basic Color Terms: Their Universality and Evolution. Berkeley, CA:
University of California Press.

Butler, Judith. 1988. “Performative Acts and Gender Constitution: An Essay in Phenomenology and Feminist
Theory.” Theatre Journal 40 (4): 519–31.

Chun, Elaine W., and Adrienne Lo. 2015. “Language and Racialization.” In The Routledge Handbook of
Linguistic Anthropology, edited by Nancy Bonvillain. Abingdon, UK: Routledge.

Clarke, Esther, Ulrich H. Reichard, and Klaus Zuberbühler. 2006. “The Syntax and Meaning of Wild Gibbon
Songs.” PLoS ONE 1 (1): e73. https://doi.org/10.1371/journal.pone.0000073.

Duranti, Alessandro. 1997. Linguistic Anthropology. Cambridge, UK: Cambridge University Press.

Epps, Patience, and Danilo Paiva Ramos. 2020. “Enactive Esthetics: The Poetics of Hup Incantation.” Journal of
Linguistic Anthropology 30 (2): 233–257.

Feldman, Orna. 2001. “Inspired by a Dream: Jessie ‘Little Doe’ Fermino, a Mashpee Indian, Is Working to
Revive the Wampanoag Language.” Spectrum. http://spectrum.mit.edu/spring-2001/inspired-by-a-dream/

Frisch, Karl von. 1993. The Dance Language and Orientation of Bees. Cambridge, MA: Harvard University
Press.
208 6 • Bibliography

Gibson, Kathleen, and Maggie Tallerman. 2011. “Introduction.” In The Oxford Handbook of Language
Evolution, edited by Kathleen Gibson and Maggie Tallerman. Oxford, UK: Oxford University Press.

Hall, Kira. 1995. “Hijra/hijrin: Language and Gender Identity.” PhD diss. University of California, Berkeley.
ProQuest Dissertations Publishing.

Haynie, Hannah, and Claire Bowern. 2016. “Phylogenetic Approach to the Evolution of Color Term Systems.”
Proceedings of the National Academy of Sciences 113 (48): 13666–71.

Herrick, J. W. 1995. Iroquois Medical Botany. Syracuse, NY: Syracuse University Press.

Hill, Jane H. 2008. The Everyday Language of White Racism. Malden, MA: Wiley-Blackwell.

Hinton, Leanne. 2011. “Revitalization of Endangered Languages.” In The Cambridge Handbook of Endangered
Languages, edited by P. K. Austin and J. Sallabank, 291–311. Cambridge, UK: Cambridge University Press.

Hinton, Leanne. 2018. “Approaches to and Strategies for Language Revitalization.” In The Oxford Handbook of
Endangered Languages, edited by Kenneth L. Rehg and Lyle Campbell. New York: Oxford University Press.

Hyde, Janet. 2005. “The Gender Similarities Hypothesis.” American Psychologist 60 (6): 581–92.

Kay, Paul, and Willett Kempton. 1984. “What Is the Sapir-Whorf Hypothesis?” American Anthropologist, 86 (1):
65–79.

Kay, Paul. 2015. “Linguistics of Color Terms.” In International Encyclopedia of the Social & Behavioral
Sciences, edited by James D. Wright, 231–4. Amsterdam, NY: Elsevier.

Keenan [Ochs], Elinor. 1974. “Norm-Makers, Norm-Breakers: Uses of Speech by Men and Women in a Malagasy
Community.” In Ethnography of Communication, edited by R. Bauman and J. Sherzer, 125–43. Cambridge,
UK: Cambridge University Press.

King, Jeanette. 2018. “Māori: Revitalization of an Endangered Language.” In The Oxford Handbook of
Endangered Languages, edited by Kenneth L. Rehg and Lyle Campbell. New York: Oxford University Press.

Kulick, Don. 1992. Language Shift and Cultural Reproduction: Socialization, Self, and Syncretism in a Papua
New Guinean Village. Cambridge, UK: Cambridge University Press.

Labov, William. 1972a. “Academic Ignorance and Black Intelligence.” June 1972. https://www.ling.upenn.edu/
courses/ling001/Labov1972.pdf.

Labov, William. 1972b. “The Logic of Nonstandard English.” In Language in the Inner City: Studies in the Black
English Vernacular, 201–40. Philadelphia, PA: University of Pennsylvania Press.

Labov, William, Ash, Sharon, Boberg, Charles. 2006. Atlas of American English: Phonetics, Phonology and
Sound Change. New York. Mouton de Gruyter.

Lakoff, George, and Mark Johnson. 1980. Metaphors We Live By. Chicago: University of Chicago Press.

Lakoff, Robin. 1975. Language and Woman’s Place. New York: Harper and Row.

Levinson, Steven. 2003. Space in Language and Cognition: Explorations in Cognitive Diversity. Cambridge, UK:
Cambridge University Press.

Lim, Yen Ying, and Peter Snyder. 2015. “Human Language: Evolutionary Precursors.” In International
Encyclopedia of the Social & Behavioral Sciences, edited by James D. Wright, 329–34. Amsterdam, NY:
Elsevier.

Lippi-Green, Rosina. 2012. English with an Accent: Language, Ideology, and Discrimination in the United
States. London and New York: Routledge.

Lucy, John. 2015. “Sapir-Whorf Hypothesis.” In International Encyclopedia of the Social & Behavioral Sciences,
edited by James D. Wright, 903–6. Amsterdam, NY: Elsevier.

Martin, Laura. 1986. “‘Eskimo Words for Snow’: A Case Study in the Genesis and Decay of an Anthropological

Access for free at openstax.org


6 • Bibliography 209

Example.” American Anthropologist, 88 (2): 418–23.

Massachusetts Institute of Technology. “Did Humans Speak through Cave Art? Ancient Drawings and
Language's Origins.” ScienceDaily. February 21, 2018. https://www.sciencedaily.com/releases/2018/02/
180221122923.htm.

Mensah, Adinkrah. 2020. “'If You Die a Bad Death, We Give You a Bad Burial’: Mortuary Practices and 'Bad
Death’ among the Akan in Ghana.” Death Studies. doi: 10.1080/07481187.2020.1762264.

Mertz, Elizabeth. 2007. The Language of Law School: Learning to “Think Like a Lawyer.” Oxford, UK: Oxford
University Press.

Miyagawa, Shigeru, Cora Lesure, and Vitor A. Nóbrega. 2018. “Cross-Modality Information Transfer: A
Hypothesis about the Relationship among Prehistoric Cave Paintings, Symbolic Thinking, and the
Emergence of Language.” Frontiers in Psychology. February 20, 2018 https://doi.org/10.3389/
fpsyg.2018.00115.

Muehlmann, S. 2014. “The Speech Community and Beyond: Language and the Nature of the Social Aggregate.”
In The Cambridge Handbook of Linguistic Anthropology, edited by N. Enfield, P. Kockelman, and J. Sidnell,
577–98. Cambridge, UK: Cambridge University Press.

Ochs, Elinor, and Schieffelin, Bambi. (1984) 2001. “Language Acquisition and Socialization: Three
Developmental Stories and Their Implications.” In Linguistic Anthropology: A Reader, edited by A. Duranti,
263–301. Oxford, UK: Blackwell.

Patterson, Francine, and Eugene Linden. 1981. The Education of Koko. New York: Holt, Rinehart, and Winston.

Purnell, Thomas, William Idsardi, and John Baugh. 1999. “Perceptual and Phonetic Experiments on American
English Dialect Identification.” Journal of Language and Social Psychology 18 (1): 10–30.

Seeley, Thomas D. 2010. Honeybee Democracy. Princeton, NJ: Princeton University Press.

Shatwell, Justin. 2012. “The Long-Dead Native Language Wopânâak Is Revived.” Yankee Magazine. October 9,
2012. https://newengland.com/yankee-magazine/living/profiles/wampanoag-language.

Solnit, Rebecca. 2008. “Men Explain Things to Me; Facts Didn't Get in Their Way.” Common Dreams. April 13,
2008. https://www.commondreams.org/views/2008/04/13/men-explain-things-me-facts-didnt-get-their-
way.

Sukiennik, Greg. 2001. “Woman Brings Tribe’s Dead Language Back to Life.” Los Angeles Times. March 24,
2001. https://www.latimes.com/archives/la-xpm-2001-mar-24-mn-42039-story.html

Sweet, Nicholas. 2019. “Ritual Contingency: Teasing and the Politics of Participation.” Journal of Linguistic
Anthropology 30 (1): 86–102.

Tannen, Deborah. 1990. You Just Don’t Understand: Women and Men in Conversation. New York: Ballantine
Books.

West, C. 1998. “When the Doctor Is a Lady: Power, Status, and Gender in Physician-Patient Encounters.” In
Language and Gender: A Reader, edited by J. Coates, 396–412. Oxford, UK: Blackwell.

“Words We’re Watching: Mansplaining.” 2018. Merriam-Webster.com Dictionary. Merriam-Webster.


https://www.merriam-webster.com/words-at-play/mansplaining-definition-history.

Yu, Ning. 2016. “Spatial Metaphors for Morality: A Perspective from Chinese.” Metaphor and Symbol 31 (2):
108–25.

Zimmerman, D., and C. West. 1975. “Sex-Roles, Interruptions, and Silences in Conversation.” In Language,
Gender and Society, edited by B. Thorne, C. Kramarae, and N. Henley, 89–101. Rowley, MA: Newbury House.
210 6 • Bibliography

Access for free at openstax.org

You might also like