#Sales Offer!| Get upto 25% Off:

ESSENTIALS OF PHYSICAL ANTHROPOLOGY

 

 

 

Bn W. W. NORTON & COMPANY NEW  YORK • LONDON

ESSENTIALS OF PHYSICAL ANTHROPOLOGY D I S C O V E R I N G O U R O R I G I N S

CLARK SPENCER LARSEN T H E O H I O S T A T E U N I V E R S I T Y

T H I R D   E D I T I O N

 

 

W. W. Norton & Company has been independent since its founding in 1923, when William Warder Norton and Mary D. Herter Norton first published lectures delivered at the People’s Institute, the adult education divi- sion of New York City’s Cooper Union. The firm soon expanded its program beyond the Institute, publishing books by celebrated academics from America and abroad. By mid century, the two major pillars of Norton’s publishing program— trade books and college texts— were firmly established. In the 1950s, the Norton family transferred control of the company to its employees, and today— with a staff of four hundred and a compara- ble number of trade, college, and professional titles published each year— W. W. Norton & Company stands as the largest and oldest publishing house owned wholly by its employees.

Copyright © 2016, 2013, 2010 by W. W. Norton & Company, Inc.

All rights reserved. Printed in the United States of America.

Editor: Eric Svendsen Editorial Assistant: Rachel Goodman Development Editor: Sunny Hwang Manuscript Editor: Andrew Pachuta Project Editor: Diane Cipollone Marketing Manager, Anthropology: Jake Schindel Media Editor: Tacy Quinn Assistant Editor, Emedia: Marina Rozova Associate Director of Production: Benjamin Reynolds Photo Editor: Stephanie Romeo Permissions Manager: Megan Jackson Permissions Specialist: Bethany Salminen Text Designer: Jillian Burr Art Director: Rubina Yeh Composition: Jouve Illustrations by Imagineering Manufacturing: R. R. Donnelley & Sons—Kendallville IN

Library of Congress Cataloging- in- Publication Data

Larsen, Clark Spencer. Essentials of physical anthropology : discovering our origins / Clark Spencer Larsen, The Ohio State University.—Third edition. pages cm Includes index. ISBN 978-0-393-93866-1 (pbk.) 1. Physical anthropology. I. Title. GN50.4.L367 2015 599.9—dc23 2015023645

W. W. Norton & Company, Inc., 500 Fifth Avenue, New York, N.Y. 10110-0017 www.wwnorton.com W. W. Norton & Company Ltd., Castle House, 75/76 Wells Street, London W1T 3QT

1 2 3 4 5 6 7 8 9 0

 

 

v

TO CHRIS AND SPENCER, WITH MY DEEPEST THANKS FOR THEIR HELP, ENCOURAGEMENT, AND

(UNWAVERING) PATIENCE

 

 

vi

CLARK SPENCER LARSEN heads the Department of Anthropology at The Ohio State University, Columbus. A native of Nebraska, he received his B.A. from Kansas State University and M.A. and Ph.D. from the Uni- versity of Michigan. Clark’s research is in bioarchaeology, skeletal biol- ogy, and paleoanthropology. He has worked in North America, Europe, and Asia. He has taught at the University of Massachusetts, Northern Illi- nois University, Purdue University, and the University of North Carolina. Since 2001, he has been a member of the faculty at Ohio State, where he is Distinguished Professor of Social and Behavioral Sciences. He teaches introductory physical anthropology, osteology, bioarchaeology, and paleoanthropology. Clark has served as president of the American Association of Physical Anthropologists and as editor- in- chief of the American Journal of Physical Anthropology. In addition to Our Origins, he has authored or edited 30 books and monographs, including Bioar- chaeology: Interpreting Behavior from the Human Skeleton, Skeletons in Our Closet, Advances in Dental Anthropology, and A Companion to Biological Anthropology.

ABOUT THE AUTHOR

 

 

vii

To the Instructor xx To the Student xxviii

CHAPTER 1 What Is Physical Anthropology? 2

PART I The Present: Foundation for the Past 19

CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory 20

CHAPTER 3 Genetics: Reproducing Life and Producing Variation 42

CHAPTER 4 Genes and Their Evolution: Population Genetics 70

CHAPTER 5 Biology in the Present: Living People 100

CHAPTER 6 Biology in the Present: The Other Living Primates 132

CHAPTER 7 Primate Sociality, Social Behavior, and Culture 164

PART II The Past: Evidence for the Present 183

CHAPTER 8 Fossils and Their Place in Time and Nature 184

CHAPTER 9 Primate Origins and Evolution: The First 50 Million Years 216

CHAPTER 10 Early Hominin Origins and Evolution: The Roots of Humanity 244

CHAPTER 11 The Origins and Evolution of Early Homo 282

CHAPTER 12 The Origins, Evolution, and Dispersal of Modern People 306

CHAPTER 13 Our Last 10,000 Years: Agriculture, Population, Biology 350

BASIC TABLE OF CONTENTS

 

 

viii

 

 

TABLE OF CONTENTS

Two-Page Spreads xix

To the Instructor xx

Tools for Teaching and Learning xxiii

Who Helped xxv

To the Student xxviii

CHAPTER 1 WHAT IS PHYSICAL ANTHROPOLOGY? 2

Big Questions 3 What Is Anthropology? 5 What Is Physical Anthropology? 7

What Do Physical Anthropologists Do? 7 What Makes Humans So Different from Other Animals?: The Six Steps to

Humanness 8 How We Know What We Know: The Scientific Method 14 Answering the Big Questions 16 Key Terms 17 Evolution Review 17 Additional Readings 17

ix

 

 

x Table of Contentsx Table of Contents

PART I THE PRESENT: FOUNDATION FOR THE PAST 19

CHAPTER 2 EVOLUTION: CONSTRUCTING A FUNDAMENTAL SCIENTIFIC THEORY 20

Big Questions 21 The Theory of Evolution: The Context for Darwin 23

Geology: Reconstructing Earth’s Dynamic History 24 Paleontology: Reconstructing the History of Life on Earth 25 Taxonomy and Systematics: Classifying Living Organisms and Identifying Their

Biological Relationships 26 Concept Check Pre-Darwinian Theory and Ideas: Groundwork for

Evolution 27 Demography: Influences on Population Size and Competition for Limited

Resources 28 Evolutionary Biology: Explaining the Transformation of Earlier Life-Forms into

Later Life-Forms 28 Concept Check Darwin Borrows from Malthus 30 The Theory of Evolution: Darwin’s Contribution 31 Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the

Discovery of DNA 33 Mechanisms of Inheritance 33 The Evolutionary Synthesis, the Study of Populations, and the Causes of

Evolution 36 DNA: Discovery of the Molecular Basis of Evolution 37

Answering the Big Questions 39 Key Terms 39 Evolution Review: Past, Present, and Future of a Fundamental Scientific

Theory 40 Additional Readings 41

CHAPTER 3 GENETICS: REPRODUCING LIFE AND PRODUCING VARIATION 42

Big Questions 43 The Cell: Its Role in Reproducing Life and Producing Variation 44 The DNA Molecule: The Genetic Code 46

DNA: The Blueprint of Life 48 The DNA Molecule: Replicating the Code 48 How Do We Know? Ancient DNA Opens New Windows on the Past 50 Concept Check The Two Steps of DNA Replication 51

Chromosome Types 51 Mitosis: Production of Identical Somatic Cells 52 Meiosis: Production of Gametes (Sex Cells) 54 Producing Proteins: The Other Function of DNA 56

 

 

Table of Contents xi

Concept Check The Two Steps of Protein Synthesis 60 Genes: Structural and Regulatory 61 Polymorphisms: Variations in Specific Genes 61

Genotypes and Phenotypes: Genes and Their Expression 63 The Complexity of Genetics 65 Answering the Big Questions 67 Key Terms 68 Evolution Review: Insights from Genetics 68 Additional Readings 69

CHAPTER 4 GENES AND THEIR EVOLUTION: POPULATION GENETICS 70

Big Questions 71 Demes, Reproductive Isolation, and Species 72 Hardy-Weinberg Law: Testing the Conditions of Genetic Equilibrium 76 Mutation: The Only Source of New Alleles 77 Natural Selection: Advantageous Characteristics, Survival, and

Reproduction 80 Patterns of Natural Selection 81 Natural Selection in Animals: The Case of the Peppered Moth and Industrial

Melanism 82 Natural Selection in Humans: Abnormal Hemoglobins and Resistance to

Malaria 84 The Geography of Sickle-Cell Anemia and the Association with Malaria 86 The Biology of Sickle-Cell Anemia and Malarial Infection 87 The History of Sickle-Cell Anemia and Malaria 87 Other Hemoglobin and Enzyme Abnormalities 89

Genetic Drift: Genetic Change due to Chance 90 Founder Effect: A Special Kind of Genetic Drift 93

Gene Flow: Spread of Genes across Population Boundaries 93 Concept Check What Causes Evolution? 97 Answering the Big Questions 97 Key Terms 98 Evolution Review: The Four Forces of Evolution 99 Additional Readings 99

CHAPTER 5 BIOLOGY IN THE PRESENT: LIVING PEOPLE 100

Big Questions 101 Is Race a Valid, Biologically Meaningful Concept? 102

Brief History of the Race Concept 102 Debunking the Race Concept: Franz Boas Shows that Human Biology Is Not

Static 103

Table of Contents xi

 

 

xii Table of Contentsxii Table of Contents

So-Called Racial Traits Are Not Concordant 103 Human Variation: Geographic Clines, Not Racial Categories 103

Life History: Growth and Development 104 The Growth Cycle: Conception through Adulthood 105 Prenatal Stage: Sensitive to Environmental Stress, Predictive of Adult Health

105 Postnatal Stage: The Maturing Brain, Preparing for Adulthood 106 Adult Stage: Aging and Senescence 109 Evolution of Human Life History: Food, Sex, and Strategies for Survival and

Reproduction 111 Concept Check Life History Stages in Humans: Prenatal, Postnatal, and

Adult 111 Prolonged Childhood: Fat-Bodied Moms and Their Big-Brained Babies 112 Grandmothering: Part of Human Adaptive Success 112

Adaptation: Meeting the Challenges of Living 113 Climate Adaptation: Living on the Margins 114

Heat Stress and Thermoregulation 114 Body Shape and Adaptation to Heat Stress 114 Cold Stress and Thermoregulation 115 Solar Radiation and Skin Color 116 Solar Radiation and Vitamin D Synthesis 117 Solar Radiation and Folate Protection 118 High Altitude and Access to Oxygen 118

Concept Check Adaptation: Heat, Cold, Solar Radiation, High Altitude 119 Nutritional Adaptation: Energy, Nutrients, and Function 120

Macronutrients and Micronutrients 120 Human Nutrition Today 121 Overnutrition and the Consequences of Dietary Excess 123

Concept Check Nutritional Adaptation 126 Workload Adaptation: Skeletal Homeostasis and Function 126 Excessive Activity and Reproductive Ecology 128

Answering the Big Questions 129 Key Terms 130 Evolution Review: Human Variation Today 130 Additional Readings 131

CHAPTER 6 BIOLOGY IN THE PRESENT: THE OTHER LIVING PRIMATES 132

Big Questions 133 What Is a Primate? 135

Arboreal Adaptation—Primates Live in Trees and Are Good at It 138 Primates Have a Versatile Skeletal Structure 138 Primates Have an Enhanced Sense of Touch 140

 

 

Table of Contents xiii

Primates Have an Enhanced Sense of Vision 141 Primates Have a Reduced Reliance on Senses of Smell and Hearing 141

Concept Check What Makes Primates Good at Living in Trees? 142 Dietary Plasticity—Primates Eat a Highly Varied Diet, and Their Teeth Reflect This

Adaptive Versatility 142 Primates Have Retained Primitive Characteristics in Their Teeth 142 Primates Have a Reduced Number of Teeth 142 Primates Have Evolved Different Dental Specializations and Functional

Emphases 143 Concept Check What Gives Primates Their Dietary Flexibility? 143

Parental Investment—Primate Parents Provide Prolonged Care for Fewer but Smarter, More Socially Complex, and Longer-Lived Offspring 146

Concept Check Primate Parenting 148 What Are the Kinds of Primates? 148

The Strepsirhines 153 Concept Check Monkey or Ape? Differences Matter 154

The Haplorhines 155 Concept Check Strepsirhines and Haplorhines Differ in Their Anatomy and

Senses 161 Answering the Big Questions 162 Key Terms 162 Evolution Review: Our Closest Living Relatives 163 Additional Readings 163

CHAPTER 7 PRIMATE SOCIALITY, SOCIAL BEHAVIOR, AND CULTURE 164

Big Questions 165 Primate Societies: Diverse, Complex, Long-Lasting 166

Diversity of Primate Societies 166 Primate Social Behavior: Enhancing Survival and Reproduction 167 Primate Residence Patterns 168 Primate Reproductive Strategies: Males’ Differ from Females’ 169

Concept Check Male and Female Reproductive Strategies 170 The Other Side of Competition: Cooperation in Primates 170

Getting Food: Everybody Needs It, but the Burden Is on Mom 172 Acquiring Resources and Transmitting Knowledge: Got Culture? 173 Vocal Communication Is Fundamental Behavior in Primate Societies 175 Answering the Big Questions 181 Key Terms 181 Evolution Review: Primate Social Organization and Behavior 182 Additional Readings 182

 

 

xiv Table of Contents

PART II THE PAST: EVIDENCE FOR THE PRESENT 183

CHAPTER 8 FOSSILS AND THEIR PLACE IN TIME AND NATURE 184

Big Questions 185 Fossils: Memories of the Biological Past 188

What Are Fossils? 188 Taphonomy and Fossilization 188 Types of Fossils 188 Limitations of the Fossil Record: Representation Is Important 191

Just How Old Is the Past? 192 Time in Perspective 192 Geologic Time: Earth History 193 Relative and Numerical Age 195 Relative Methods of Dating: Which Is Older, Younger, the Same Age? 196

Stratigraphic Correlation 196 Chemical Dating 196 Biostratigraphic (Faunal) Dating 197 Cultural Dating 198

Absolute Methods of Dating: What Is the Numerical Age? 198 The Radiometric Revolution and the Dating Clock 198 The Revolution Continues: Radiopotassium Dating 203 Non-Radiometric Absolute Dating Methods 205

Genetic Dating: The Molecular Clock 207 Concept Check How Old Is It? 208 Reconstruction of Ancient Environments and Landscapes 209

The Driving Force in Shaping Environment: Temperature 210 Chemistry of Animal Remains and Ancient Soils: Windows onto Diets and

Habitats 211 Answering the Big Questions 213 Key Terms 214 Evolution Review: The Fossil Record 214 Additional Readings 215

CHAPTER 9 PRIMATE ORIGINS AND EVOLUTION: THE FIRST 50 MILLION YEARS 216

Big Questions 217 Why Did Primates Emerge? 218 The First True Primate: Visual, Tree-Dwelling, Agile, Smart 220

Primates in the Paleocene? 220 Eocene Euprimates: The First True Primates 220 The Anthropoid Ancestor: Euprimate Contenders 224 The First Anthropoids 225

 

 

Table of Contents xv

Early Anthropoids Evolve and Thrive 227 Concept Check When Were They Primates?: Anatomy through Time 227 Coming to America: Origin of New World Higher Primates 230

How Anthropoids Got to South America 230 Apes Begin in Africa and Dominate the Miocene Primate World 231 Apes Leave Africa: On to New Habitats and New Adaptations 234

Apes in Europe: The Dryopithecids 234 Apes in Asia: The Sivapithecids 235 Dead End in Ape Evolution: The Oreopithecids 235 Climate Shifts and Habitat Changes 238 Miocene Ape Survivors Give Rise to Modern Apes 238

Apes Return to Africa? 238 Concept Check The First Apes: A Remarkable Radiation 239 Monkeys on the Move 239 Answering the Big Questions 241 Key Terms 242 Evolution Review: Primate Social Organization and Behavior:

The Deep Roots of the Order Primates 242 Additional Readings 243

CHAPTER 10 EARLY HOMININ ORIGINS AND EVOLUTION: THE ROOTS OF HUMANITY 244

Big Questions 245 What Is a Hominin? 246

Bipedal Locomotion: Getting Around on Two Feet 248 Nonhoning Chewing: No Slicing, Mainly Grinding 248

Why Did Hominins Emerge? 251 Charles Darwin’s Hunting Hypothesis 251

Concept Check What Makes a Hominin a Hominin? 252 Peter Rodman and Henry McHenry’s Patchy Forest Hypothesis 254 Owen Lovejoy’s Provisioning Hypothesis 254 Sexual Dimorphism and Human Behavior 255 Bipedality Had Its Benefits and Costs: An Evolutionary Trade-Off 255

What Were the First Hominins? 256 The Pre-Australopithecines 256

Sahelanthropus tchadensis (7–6 mya) 257 Orrorin tugenensis (6 mya) 257 Ardipithecus kadabba and Ardipithecus ramidus (5.8–4.4 mya) 258

Concept Check The Pre-Australopithecines 263 The Australopithecines (4–1 mya) 264

Australopithecus anamensis (4 mya) 265 Australopithecus afarensis (3.6–3.0 mya) 266 Australopithecus (Kenyanthropus) platyops (3.5 mya) 269

 

 

xvi Table of Contents

Diversification of the Homininae: Emergence of Multiple Evolutionary Lineages from One (3–1 mya) 269

Australopithecus garhi (2.5 mya) 270 The First Tool Makers and Users: Australopithecus or Homo? 270

Evolution and Extinction of the Australopithecines 273 Concept Check The Australopithecines 276 Answering the Big Questions 280 Key Terms 280 Evolution Review: The First Hominins 281 Additional Readings 281

CHAPTER 11 THE ORIGINS AND EVOLUTION OF EARLY HOMO 282

Big Questions 283 Homo habilis: The First Species of the Genus Homo 285

The Path to Humanness: Bigger Brains, Tool Use, and Adaptive Flexibility 285

Homo habilis and Australopithecus: Similar in Body Plan 287 Homo habilis’s Adaptation: Intelligence and Tool Use Become Important 287 Habitat Changes and Increasing Adaptive Flexibility 288

Concept Check Homo habilis: The First Member of Our Lineage 288 Homo erectus: Early Homo Goes Global 289

Homo erectus in Africa (1.8–.3 mya) 290 Homo erectus in Asia (1.8–.3 mya) 293 Homo erectus in Europe (1.2 million–400,000 yBP) 296 Evolution of Homo erectus: Biological Change, Adaptation, and Improved

Nutrition 297 Patterns of Evolution in Homo erectus 302

Concept Check Homo erectus: Beginning Globalization 303 Answering the Big Questions 304 Key Terms 305 Evolution Review: The Origins of Homo 305 Additional Readings 305

CHAPTER 12 THE ORIGINS, EVOLUTION, AND DISPERSAL OF MODERN PEOPLE 306

Big Questions 307 What Is So Modern about Modern Humans? 309 Modern Homo sapiens: Single Origin and Global Dispersal or Regional

Continuity? 309 What Do Homo sapiens Fossils Tell Us about Modern Human Origins? 311

Early Archaic Homo sapiens 311 Archaic Homo sapiens in Africa (350,000–200,000 yBP) 312

 

 

Table of Contents xvii

Early Archaic Homo sapiens in Asia (350,000–130,000 yBP) 312 Early Archaic Homo sapiens in Europe (500,000–130,000 yBP) 313 Early Archaic Homo sapiens’ Dietary Adaptations 313

Late Archaic Homo sapiens 314 Late Archaic Homo sapiens in Asia (60,000–40,000 yBP) 315 Late Archaic Homo sapiens in Europe (130,000–30,000 yBP) 316 The Neandertal Body Plan: Aberrant or Adapted? 319 Neandertal Hunting: Inefficient or Successful? 321 Neandertals Buried Their Dead 324 Neandertals Talked 325 Neandertals Used Symbols 327

Early Modern Homo sapiens 327 Concept Check Archaic Homo sapiens 328

Early Modern Homo sapiens in Africa (200,000–6,000 yBP) 329 Early Modern Homo sapiens in Asia (90,000–18,000 yBP) 331 Early Modern Homo sapiens in Europe (35,000–15,000 yBP) 332

Modern Behavioral and Cultural Transitions 334 How Has the Biological Variation in Fossil Homo sapiens Been

Interpreted? 335 Ancient DNA: Interbreeding between Neandertals and Early Modern People? 336

Concept Check Early Modern Homo sapiens 337 Living People’s Genetic Record: Settling the Debate on Modern Human Origins 338

Assimilation Model for Modern Human Variation: Neandertals Are Still with Us 339

Concept Check Models for Explaining Modern Homo sapiens’ Origins 340 Modern Humans’ Other Migrations: Colonization of Australia, the Pacific, and

the Americas 340 Down Under and Beyond: The Australian and Pacific Migrations 342 Arrival in the Western Hemisphere: The First Americans 344

Answering the Big Questions 348 Key Terms 349 Evolution Review: The Origins of Modern People 349 Additional Readings 349

CHAPTER 13 OUR LAST 10,000 YEARS: AGRICULTURE, POPULATION, BIOLOGY 350

Big Questions 351 The Agricultural Revolution: New Foods and New Adaptations 353

Population Pressure 354 Regional Variation 355 Survival and Growth 359

Agriculture: An Adaptive Trade-Off 360 Population Growth 360 Environmental Degradation 361

Concept Check The Good and Bad of Agriculture 362

 

 

xviii Table of Contents

How Did Agriculture Affect Human Biology? 362 The Changing Face of Humanity 363

Two Hypotheses 363 Implications for Teeth 365

Concept Check Soft Food and Biological Change 365 Building a New Physique: Agriculture’s Changes to Workload/Activity 366 Health and the Agricultural Revolution 369

Population Crowding and Infectious Disease 369 Concept Check Labor, Lifestyle, and Adaptation in the Skeleton 370

The Consequences of Declining Nutrition: Tooth Decay 371 Nutritional Consequences Due to Missing Nutrients: Reduced Growth and

Abnormal Development 371 Nutritional Consequences of Iron Deficiency 373

Concept Check Health Costs of Agriculture 374 Nutritional Consequences: Heights on the Decline 375

If It Is So Bad for You, Why Farm? 375 The Past Is Our Future 375 Our Ongoing Evolution 376 Answering the Big Questions 378 Key Terms 379 Evolution Review: Setting the Stage for the Present and Future 379 Additional Readings 380

Appendix: The Skeleton A1

Glossary A11

Glossary of Place Names A19

Bibliography A21

Permissions Acknowledgments A47

Index A51

 

 

xix

T WO- PAGE SPRE ADS

I1

ENHANCED TOUCH

Primates have an enhanced sense of touch. This sensitivity is due in part to the presence of dermal ridges (fingerprints and toe prints) on the inside surfaces of the hands and feet. The potto, a prosimian, has primitive dermal ridges, whereas the human, a higher primate, has more derived ridges, which provide better gripping ability.

Em er

gi ng

c an

op y

M ai

n ca

no py

U nd

er st

or y

GENERALIZED SKELETAL STRUCTURE

Primates have a generalized skeletal structure. The bones that make up the shoulders, upper limbs, lower limbs, and other major joints such as the hands and feet are separate, giving primates a great deal of flexibility when moving in trees. In this monkey skeleton, note the grasping hands and feet, the long tail, and the equal length of the front and hind limbs relative to each other.

REDUCED SMELL

Primates have a reduced sense of smell. The smaller and less projecting snouts of most primates indicate their decreased reliance on smell.

DIETARY VERSATILITY

Primates have dietary plasticity. Part of the record of primate dietary adaptation is found in the teeth. The red colobus monkey dentition shown here is typical of a catarrhine dentition with a 2/1/2/3 dental formula. Note the differences in morphology of the four different tooth types: incisors (I1, I2), canines (C), premolars (P3, P4), and molars (M1, M2, M3).

ENHANCED VISION

Primates have an enhanced sense of vision. Evolution has given primates better vision, including increased depth perception and seeing in color. The eyes’ convergence provides significant overlap in the visual fields and thus greater sense of depth.

Human Potto

Overlapping visual fields

Taï Forest

MonkeyDog

Reduced snout length

I1 I2I2

CC

P3P3 P4P4 M1

M2

M3

M1

M2

M3

I1I1 I2I2 CC

P3P3

P4P4

M1

M2

M3

M1

M2

M3

Black-and-white colobus

Campbell’s

Chimpanzee

Demidoff’s galago

Diana monkey

Human

Lesser spot-nosed

Olive colobus

Potto

Putty-nosed

Red colobus

Sooty mangabey

Thomas’s galago

Eagle

F I G U R E

6.2 Primate Adaptation in Microcosm: The Taï Forest, Ivory Coast, West Africa

Apes Leave Africa: On to New Habitats and New Adaptations | 237236 | CHAPTER 9 Primate Origins and Evolution: The First 50 Million Years

Primate evolution began with primitive primates in the Eocene, setting the stage for the origin of all hominoids. Euprimates of the Eocene had the basic characteristics of living primates, such as convergent eye orbits and grasping digits. In the last 20 million years, primates diversified in appearance and behavior. These changes included the shift, for some, from life in the trees to life on the ground, and eventually the beginning of bipedality in the late miocene. (Based on Fleagle, J. G. Primate Adaptation and Evolution, 2nd ed. 1999. Academic Press.)

Scenes from the late Eocene in the Paris Basin. Top: The diurnal Adapis is feeding on leaves. Bottom: Several taxa of omomyids (Pseudoloris, Necrolemur, Microchoerus). Note the large eyes, a nocturnal adaptation, typical of both ancient and modern prosimians who are active at night.

Scene from the early Miocene of Rusinga Island, Kenya. Apes first appeared during this period, and these are the first apes (two species of Proconsul, Dendropithecus, Limnopithecus). These and other taxa form the ancestry of all later apes and hominins. Note the range of habitats occupied by these primates within the forest, including some in the middle and lower canopies and some on the forest floor. These primates show a combination of monkeylike and apelike features, in the skeleton and skull, respectively.

Scenes from the early Oligocene of the Fayum, Egypt. These anthropoid ancestors include Aegyptopithecus, Propliopithecus, and Apidium. These primates were adept arborealists, using their hands and feet for climbing and feeding.

Convergent eyes and grasping hands

Large eyes for nocturnal vision

Eocene 34–56 mya

Oligocene 23–34 mya Miocene 5.3–23 mya

Quadrupedal, monkeylike primate with superb arboreal skills

Quadrupedal, apelike primate. Note the lack of a tail, an ape characteristic.

Eocene-Oligocene-Miocene Habitats and Their Primates

F I G U R E

9.21

Figure 1.3 The Six Big Events of Human Evolution: Bipedalism, Nonhoning Chewing, Dependence on Material Culture, Speech, Hunting, and Domestication of Plants and Animals pp. 10–11

Figure 3.17 Protein Synthesis pp. 58–59

Figure 6.2 Primate Adaptation in Microcosm: The Taï Forest, Ivory Coast, West Africa pp. 136–137

Figure 9.21 Eocene– Oligocene– Miocene Habitats and Their Primates pp. 236–237

Figure 10.16 From Discovery to Understanding: Ardipithecus of Aramis pp. 260–261

 

 

xx

TO THE INSTRUCTOR

HOW THIS BOOK CAN HELP YOUR STUDENTS DISCOVER PHYSICAL ANTHROPOLOGY

IT IS ABOUT ENGAGEMENT

Teaching is about engagement— connecting the student with knowledge, making it real to the student, and having the student come away from the course with an understanding of core concepts. Essentials of Physical Anthropology seeks to engage the student in the learning process. Engaging the student is perhaps more of a challenge in the study of phys- ical anthropology than in the study of other sciences, mainly because the student has likely never heard of the subject. The average student has probably taken a precollege course in chemistry, physics, or biology. Physical anthropology, though, is rarely mentioned or taught in precollege settings. Commonly, the student first finds out about the subject when an academic advisor explains that physical anthro- pology is a popular course that fulfills the college’s natural science requirement.

Once taking the course, however, that same student usually connects quickly with the subject because so many of the topics are familiar— fossils, evolution, race, genet- ics, DNA, monkeys, forensic investigations, and origins of speech, to name a few. The student simply had not real- ized that these separately engaging topics come under the umbrella of one discipline, the subject of which is the study of human evolution and human variability.

Perhaps drawn to physical anthropology because it focuses on our past and our present as a species, the student quickly sees the fundamental importance of the discipline. In Discover magazine’s 100 top stories of 2009, 18 were from physical anthropology. Three topics from the field were in the top 10, including the remarkable new discovery of our earliest human ancestor, Ardipithecus. So important was this discovery that Science, the leading international professional science journal, called it the “Breakthrough of the Year” for

2009. The discussions in this textbook of topics familiar and unfamiliar give the student stepping- stones to science and to the centrality of physical anthropology as a window into understanding our world. Whether the students find the material familiar or unfamiliar, they will see that the book relates the discipline to human life: real concerns about human bodies and human identity. They will see themselves from an entirely different point of view and gain new awareness.

In writing this book, I made no assumptions about what the reader knows, except to assume that the reader— the stu- dent attending your physical anthropology class— has very little or no background in physical anthropology. As I wrote the book, I constantly reflected on the core concepts of phys- ical anthropology and how to make them understandable. I combined this quest for both accuracy and clarity with my philosophy of teaching— namely, engage the student to help the student learn. Simply, teaching is about engagement. While most students in an introductory physical anthro- pology class do not intend to become professional physical anthropologists, some of these students become interested enough to take more courses. So this book is written for stu- dents who will not continue their study of physical anthro- pology, those who get “hooked” by this fascinating subject (a common occurrence!), and those who now or eventually decide to become professionals in the field.

The book is unified by the subject of physical anthropol- ogy. But equally important is the central theme of science— what it is, how it is done, and how scientists (in our case, anthropologists) learn about the natural world. I wrote the book so as to create a picture of who humans are as organ- isms, how we got to where we are over the last millions of years of evolution, and where we are going in the future in light of current conditions. In regard to physical anthro- pology, the student should finish the book understanding human evolution and how it is studied, how the present helps us understand the past, the diversity of organisms living and past, and the nature of biological change over

 

 

xxi

time and across geography. Such knowledge should help the student answer questions about the world. For example, how did primates emerge as a unique group of mammals? Why do people look different from place to place around the world? Why is it important to gain exposure to sunlight yet unsafe to prolong that exposure? Why is it unhealthy to be excessively overweight? Throughout their history, what have humans eaten, and why is it important to know?

I have presented such topics so that the student can come to understand the central concepts and build from them a fuller understanding of physical anthropology. Throughout the book, I emphasize hypothesis testing, the core of the scientific method, and focus on that process and the excite- ment of discovery. The narrative style is personalized. Often I draw on my own experiences and those of scientists I know or am familiar with through their teaching and writing, to show the student how problems are addressed through field- work or through laboratory investigations.

Scientists do not just collect facts. Rather, they collect data and make observations that help them answer questions about the complex natural world we all inhabit. Reflecting this practice, Essentials of Physical Anthropology is a collec- tion not of facts for the student to learn but of answers to questions that help all of us understand who we are as living organisms and our place in the world. Science is a way of knowing, it is a learning process, and it connects our lives with our world. In these ways, it is liberating.

HOW THE BOOK IS ORGANIZED

The book is divided into two parts. Following an introduc- tory overview of anthropology and physical anthropology, Part  I presents the key principles and concepts in biology, especially from an evolutionary perspective. This material draws largely on the study of living organisms, including humans and nonhuman primates. Because much of our understanding of the past is drawn from what we have learned from the present, this part lays the foundation for the presentation in Part II— the past record of primate and human evolution. In putting the record of the living up front, this book departs from the style of most other introductory physical anthropology textbooks, which start out with the earliest record and end with the living. This book takes the position that most of what we learn about the past is based on theory and principles learned from the living record. Just as all of Charles Darwin’s ideas were first derived from seeing living plants and animals, much of our understanding of function and adaptation comes from living organisms as models. Therefore, this book views the living as the window

into what came before— the present contextualizes and informs our understanding of the past. It is no mistake, then, that Discovering Our Origins is the subtitle of the book. The origins of who we are today do not just lie in the record of the past, but are very much embodied in the living. Our origins are expressed in our physical makeup (bone, teeth, and muscles), in our behavior, and in so many other ways that the student taking this course will learn about from this book and from you. You can teach individual chapters in any order, and that is partly because each chapter reinforces the central point: we understand our past via what we see in the living.

Part II presents evidence of the past, covering more than 50  million years of primate and human evolution. Most textbooks of this kind end the record of human evolution at about 25,000 years ago, when modern Homo sapiens evolved worldwide. This textbook also provides the record since the appearance of modern humans, showing that important bio- logical changes occurred in just the last 10,000 years, largely relating to the shift from hunting and gathering to the domestication of plants and animals. Food production was a revolutionary development in the human story, and Part II presents this remarkable record, including changes in health and well- being that continue today. A new subdiscipline of physical anthropology, bioarchaeology, is contributing pro- found insights into the last 10,000 years, one of the most dynamic periods of human evolution.

During this period, a fundamental change occurred in how humans obtained food. This change set the stage for our current environmental disruptions and modern living conditions, including global warming, the alarming global increase in obesity, and the rise of health threats such as newly emerging infectious diseases, of which there is little under- standing and for which scientists are far from finding cures.

CHANGES IN THE THIRD EDITION

Reflecting the dynamic nature of physical anthropology, there are numerous revisions and updates throughout this new, third edition of Essentials of Physical Anthropology. These updates provide content on the cutting- edge developments in the discipline, give new ways of looking at older findings, and keep the book engaging and timely for both you and your students. Although the core principle of the book remains the same, namely the focus on evolution, the revi- sions throughout the book present new insights, new discov- eries, and new perspectives. Other changes are intended to give added focus and clarity and to increase the visual appeal that supports the pedagogy of engagement and learning:

Instructor xxi

 

 

• New content on biocultural adaptation. Anthropol- ogists provide important insights into how humans’ remarkable intelligence is related to their evolution- ary success. This third edition presents new research on the role of social learning and the retention of knowledge— the accumulation of information— over many generations.

• New primate taxonomy. In order to inform students about the latest developments in primate classifica- tion, the third edition has shifted from the tradi- tional, grade- based approach used in the previous editions to the cladistics, or phylogenetic, approach. This approach provides students with a classifica- tion based on ancestor- descendant evolutionary relationships.

• New content on developments in genetics that are altering our understanding of phenotype. We are learning that non- protein coding DNA, often considered “junk” DNA, has important implications for various other instructions in the genome. Similarly, the rapidly expanding field of epigenetics is revealing evolution- ary change without alteration of DNA.

• New content on race and human variation in Chapter 5. • New content on maladaptive human behavior and health

outcomes such as obesity. The role of environment is fundamental in understanding patterns of health in very recent human evolution, including the impacts of the creation of obesogenic environments, the alarming rise in obesity globally, and the causes and consequences of these changing circumstances and outcomes.

• New content on fossil primate and hominin discoveries. Exciting new discoveries in early primate evolution from Africa and Asia are revealing the enormous variety and complexity of species. New discoveries from East Africa reveal that although all australo- pithecines were bipedal, some retained arboreal behavior relatively late in the evolution of these early hominins. New discovery of stone tools dat- ing to 3.3 million years ago—700,000 years earlier than previously known—from East Africa shows the beginnings of humankind’s reliance on material culture. Once thought to be the domain of Homo, these early dates show use of tools by earlier aus- tralopithecines, long before the origins of our genus. These discoveries continue to illustrate the com- plexity of early hominin evolution. New evidence from chemical and tooth wear analyses reveals that at least some later australopithecines were eating significant quantities of low- quality vegetation,

including grasses on the African savanna, confirming the long- held notion that some had highly specialized diets.

• New findings on the origins of cooking and its importance in human evolution. Controlled use of fire dates to as early as 1 mya in South Africa. This innovation provided a means for cooking meats and starches, thereby increasing the digestibility of these foods. New research suggests that cooking and nutri- tional changes associated with cooking may have “fueled” the increase in brain and body size in early hominins.

• New content on the appearance and evolution of modern Homo sapiens and the Neandertal genome. Analysis of the direction and pattern of scratches on the incisors of Neandertals reveals that they were pre- dominantly right- handed. In addition to showing this modern characteristic, this finding reveals that this earlier form of H. sapiens had brain laterality, a feature linked to speech. Neandertals talked. New genetic evidence reveals the presence of Neander- tal genes in modern humans, consistent with the hypothesis that modern H. sapiens interbred with Neandertals. Newly discovered hominin fossils from Denisova, Siberia, dating to the late Pleistocene represent a genome that is different from Neander- tals’ and modern H. sapiens’. This newly discovered “Denisovan” genome is also found in people living today in East Asia, suggesting that modern H. sapiens encountered Neandertals as well as other populations once in Europe.

• New findings on the future of humankind. The study of melting ice caps and glaciers around the world today reveals a dramatic warming trend. As temperatures rise, habitats are in the process of changing. These environmental changes will provide a context for evolution, both in plants and in animals. These fac- tors, coupled with reduction in species diversity, are creating new health challenges for humans today and for the foreseeable future.

• Revision of content to enhance clarity. I have contin- ued to focus on helping students understand core concepts, with considerable attention given to cell biology, genetics, DNA, race and human variation, primate taxonomy, locomotion, and dating methods. As in previous editions, I paid careful attention to the clarity of figure captions. The captions do not simply repeat text. Instead, they offer the student additional details relevant to the topic and occasional questions about concepts that the figures convey.

xxii Instructor

 

 

• Greatly enhanced art program. The new edition con- tains over 100 new or revised figures, often using a new “photorealistic” style. The book adds several full- color two- page spreads developed by Mauri- cio Antón, a world- renowned artist with expertise in conveying past life through wonderful visual presentations.

• “Evolution Review” sections. At the end of each chapter, an “Evolution Review” section summarizes material on evolution in that chapter and includes assignable questions about concepts and content. Suggested answers appear in the Instructor’s Manual.

• InQuizitive. Norton’s new formative and adaptive online learning resource improves student under- standing of the big picture concepts of physical anthropology. Students receive personalized quiz questions on the topics they need the most help with. Engaging, game- like elements motivate students as they learn. These are intended for use in teaching face- to- face, blended, or online class formats.

• New lab manual. This text now has a new lab manual, the Lab Manual for Biological Anthropology—Engaging with Human Evolution by K. Elizabeth Soluri and Sabrina C. Agarwal. This flexible and richly illus- trated manual is designed to support or enhance your current labs and collections, or work on its own. Attractively priced, discount bundles can be pur- chased including this text.

AIDS TO THE LEARNING PROCESS

Each chapter opens with a vignette telling the story of one person’s discovery that relates directly to the central theme of the chapter. This vignette is intended to draw your stu- dents into the excitement of the topic and to set the stage for the Big Questions that the chapter addresses.

BIG QUESTION learning objectives are introduced early in the chapter to help your students organize their reading and understand the topic.

CONCEPT CHECKS are scattered throughout each chap- ter and immediately follow a major section. These aids are intended to help your students briefly revisit the key points they have been reading about.

LOCATOR MAPS are placed liberally throughout the book. College- level instructors tend to hope that students have a good sense of geography, but like a lot of people who do not

look at places around the world on a daily basis, students often need reminders about geography. In recognition of this, locator maps in the book’s margins show the names and locations of places that are likely not common knowledge.

PHOTOREALISTIC ART YOU CAN “TOUCH”: Designed to give students an even better appreciation for the feel of the discipline, the art program has been substantially reworked. Now most illustrations of bones and skeletons have an almost photorealistic feel, and most primates were redrawn for a high degree of realism. This book helps your students visualize what they are reading about by including hundreds of images, many specially prepared for the book. These illustrations tell the story of physical anthropology, including key processes, central players, and important con- cepts. As much thought went into the pedagogy behind the illustration program as into the writing of the text.

DEFINITIONS are also presented in the text’s margins, giving your students ready access to what a term means generally in addition to its use in the associated text. For convenient reference, defined terms are signaled with bold- face page numbers in the index.

At the end of each chapter, ANSWERING THE BIG QUESTIONS presents a summary of the chapter’s central points organized along the lines of the Big Questions pre- sented at the beginning of the chapter.

The study of evolution is the central core concept of physical anthropology. The newly introduced EVOLUTION REVIEW section at the end of each chapter discusses topics on evolution featured in the chapter and asks questions that will help the student develop a focused understanding of content and ideas.

INQUIZITIVE is our new game- like, formative, adaptive assessment program featuring visual and conceptual ques- tions keyed to each chapter’s learning objectives from the text. InQuizitive helps you track and report on your students’ progress to make sure they are better prepared for class.

Join me now in engaging your students in the excitement of discovering physical anthropology.

TOOLS FOR TEACHING AND LEARNING

The Essentials of Physical Anthropology teaching and learning package provides instructors and students with all the tools they need to visualize anthropological concepts, learn key vocabulary, and test knowledge.

Instructor xxiii

 

 

FOR INSTRUCTORS

InQuizitive

New InQuizitive online formative and adaptive assessment is available for use with Essentials of Physical Anthropology, Third Edition, featuring interactive and engaging questions with answer- specific feedback. InQuizitive features ques- tions designed to help students better understand the core objectives of each chapter. Built to be intuitive and easy to use, InQuizitive makes it a snap to assign, assess, and report on student performance and help keep your class on track. Options are available to integrate InQuizitive into your LMS or Coursepack. Contact your local  W.  W.  Norton representative for details.

Lab Manual and Workbook for Biological Anthropology— Engaging with Human Evolution by K. Elizabeth Soluri and Sabrina C. Agarwal.

This new manual captures student interest and illustrates the discipline with the vivid images— every chapter contains large detailed figures, photographs that are properly scaled, and drawings of bones and fossils with an almost three- dimensional appearance. The labs are grouped into four units of four chapters each: 1) genetics/evolutionary theory; 2) human osteology and forensics; 3) primatology; and 4) paleoanthropology. No topic is over- or underemphasized, and the manual is flexibly designed to be used as a whole, or as individual labs, and with a school’s cast and photo collec- tion or with the sample photos provided. Each lab has unique Critical Thinking Questions to go with Chapter Review and Lab Exercises. This manual is available at student friendly prices, either as a stand- alone volume or bundled with this text, or as a custom volume.

Coursepacks

Available at no cost to professors or students, Norton Coursepacks for online or hybrid courses are available in a variety of formats, including all versions of Blackboard and WebCT. Content includes review quizzes, flash cards, and links to animations and videos. Coursepacks are available from wwnorton.com/instructors.

New Animations

These new animations of key concepts from each chapter are available in either the Coursepacks, or from wwnorton.com/ instructors. Animations are brief, easy to use, and great for explaining concepts either in class or in a distance- learning environment.

New Videos

This new streaming video service is now available through Norton Coursepacks and at wwnorton.com/instructors. These one- to seven- minute educational film clips from across the discipline but with an emphasis on paleoanthro- pology and primatology help students see and think like anthropologists and make it easy for instructors to illustrate key concepts and spark classroom discussion.

Update PowerPoint Service

To help cover what is new in the discipline, each semester we will provide a new set of supplemental lectures, notes, and assessment material covering current and breaking research. Prepared by Laurie Reitsema (University of Georgia) and with previous updates from Kathy Droesch (Suffolk County Community College), this material is available for download at wwnorton.com/instructors.

PowerPoint Slides and Art JPEGs

Designed for instant classroom use, these slides prepared by Jeremy DeSilva (Boston University) using art from the text are a great resource for your lectures. All art from the book is also available in PowerPoint and JPEG formats. Download these resources from wwnorton.com/instructors.

Instructor’s Manual

Prepared by Nancy Tatarek (Ohio University) and Greg Laden, this innovative resource provides chapter summaries, chapter outlines, lecture ideas, discussion topics, suggested reading lists for instructors and students, a guide to “Writ- ing about Anthropology,” suggested answers to Evolution Matters questions, and teaching materials for each video.

Test Bank

Prepared by Renee Garcia (Saddleback College) and Greg Laden, this Test Bank contains multiple- choice and essay questions for each chapter. It is downloadable from Norton’s Instructor’s Website and available in Word, PDF, and ExamView® Assessment Suite formats. Visit wwnorton.com/ instructors.

Ebook

An affordable and convenient alternative, Norton ebooks retain the content and design of the print book and allow students to highlight and take notes with ease, print chap- ters as needed, and search the text.

xxiv Instructor

 

 

WHO HELPED

I owe much to the many people who made this book possible, from the planning and writing of the first and sec- ond editions, and now this third edition. First and foremost, I thank my wife, Christine, and son, Spencer, who helped in innumerable ways. They were my captive audience: without protest, they listened to my ideas at the dinner table, on family trips, and in other places where we probably should have been talking about other things. Chris read many drafts of chapters and gave great advice on when and where to cut, add, or rethink. I thank my parents, the late Leon and Patricia Larsen, who introduced me to things old and sparked my interest in the human past.

Jack Repcheck first approached me about writing a text- book on introductory physical anthropology. His power of persuasion, combined with my own interest in the discipline and its presentation to college students, was instrumental in reeling me in and getting the project off the ground. Jack and others at W. W. Norton & Company made the process of writing the book a great experience in all ways, from writ- ing to publication. On the first  edition, I began work with editors John Byram and then Leo Wiegman. I am indebted to Pete Lesser, who took on the project after Leo. Pete gave direction on writing and production, provided very helpful feedback on presentation and pedagogy, and orchestrated the process of review, revision, and production— all without a hitch. Under Pete’s guidance, the first edition became the most widely used textbook in physical anthropology. Jack Repcheck continued the project in preparation for the sec- ond edition. The preparation of the third edition was over- seen by editor Eric Svendsen. His advice and guidance were central to seeing the book come to fruition. Tacy Quinn recently joined the team and has spearheaded the develop- ment of new media for this edition including InQuizitive. Marina Rozova does an excellent job developing the core supplement package for each edition. Kurt Wildermuth edited the entire manuscript for the first two editions. His skill as an editor and staying on top of content from begin- ning to end added enormously to the book’s presentation and readability. Sunny Hwang has now taken Kurt’s place and has especially helped with revisions in the end- of- chapter mate- rial and the on- line supplements program. Diane Cipollone was instrumental in producing these pages and directing a wide variety of editing issues that arose, and the entire team is now supported by Rachel Goodman. Ben Reynolds guided the process of production from beginning to end. I am also grateful to Mauricio Antón for his wonderful new illustra- tions of six “big events” of human evolution in chapter 1, the new rendition of the Taï Forest primates as a microcosm of

primate adaptation in chapter 6, and the Eocene, Oligocene, and Miocene primates and their habitats in chapter 9. Greg Laden, Renee Garcia, and Nancy Tatarek’s timely and effi- cient completion of the Test Bank and Instructor’s Manual is much appreciated. Laurie Reitsema has been recently added to the team producing our valuable update PowerPoints each semester, and I thank Kathy Droesch for her past work on these updates.

With the input of instructors and focus group attendees who are included in the reviewer list, we have created an extensive new media and assessment suite for the third  edi- tion. However, my thanks for extensive work in developing InQuizitive and our new animations go to Tracy Betsinger of SUNY Oneonta, Ashley Hurst of University of Texas at San Antonio, Kristina Killgrove of University of West Florida, Greg Laden, Joanna Lambert of the University of Colorado, and Heather Worne of University of Kentucky, with further thanks to contributors Jaime Ullinger, Quinnipiac University, and Nancy Cordell, South Puget Sound Community College. And thanks to Sandra Wheeler of University of Central Flor- ida, Ellen Miller of Wake Forest University, Bonnie Yoshida of Grossmont College, Jacqueline Eng of Western Michigan University, Jeremy DeSilva of Boston University,  K.  Eliza- beth Soluri of College of Marin, and again Nancy Cordell of South Puget Sound Community College for their important feedback and reviews of these resources.

Thanks go to former and current graduate students and faculty colleagues at the Ohio State University who helped in so many ways. I offer a very special thanks to Tracy Betsinger, who assisted in a number of aspects of the book. For the first edition, she read drafts of chapters at various stages and helped in figure selection, in glossary compilation, and as a sounding board in general for ideas that went into the book. For the second edition, she offered very helpful suggestions for revisions. Thanks to Jaime Ullinger, who provided the content and data for the box on PTC tasting. Tracy, Jaime, Jim Gosman, Dan Temple, Haagen Klaus, and Josh Sadvari read parts or all of the manuscript and offered great advice. For all three editions, I had many helpful discussions with Scott McGraw about primate behavior, evolution, and tax- onomy. Scott also provided advice on the production  of the two- page spreads on both primate diversity and eagle predation in the Taï Forest, Ivory Coast (chapters  6 and 7). For this edition, John Fleagle provided valuable support reviewing details in most of the new primate illustrations, in particular the two- page spreads, and every new piece of art was first reviewed in the larger Our  Origins volume by Arthur Durband, Andrew Kramer, and Sandra Wheeler. Doug Crews gave advice on the complexities of primate (including human) biology and  life history. Haagen Klaus

Instructor xxv

 

 

provided materials for and help on the two- page spread on the biological consequences of the agricultural revolution and many other helpful comments and suggestions for revision. Barbara Piperata advised me on key aspects of modern human biology and nutrition science, and Dawn Kitchen provided discussion and help on the fundamentals of primate communication and how best to present it. Josh Sadvari was indispensable in the creation of the Evolution Review sections at the end of each chapter.

Over the years, I have had helpful conversations with my teachers, colleagues, and students about areas of their exper- tise, and these people have influenced the development of the book in so many ways. I am grateful to Patricia  J.  O’Brien and Milford  H.  Wolpoff, my respective undergraduate and graduate advisors. Both were instrumental in developing my interest in science and the wonderful profession I work in. I thank Barry Bogin, Kristen Hawkes, Jim O’Connell, David Thomas, Bob Kelly, Jerry Milanich, Bruce Smith, Kris Gremillion, Bonnie McEwan, Matt Cartmill, Dale Hutchinson, Chris Ruff, Simon Hillson, Michael Schultz, Sam Stout, Doug Ubelaker, Dan Sellen, Clark Howell, Rick Steckel, Phil Walker, John Relethford, Mark Weiss, Mar- garet Schoeninger, Karen Rosenberg, Lynne Schepartz, Fred Smith, Brian Hemphill, Bruce Winterhalder, Meg Conkey, Desmond Clark, Erik Trinkaus, Katherine Russell, Vin Steponaitis, Mark Teaford, Richard Wrangham, Jerry Rose, Mark Cohen, William Bass, Loring Brace, Stanley Garn, Frank Livingstone, Phil Gingerich,  T.  Dale Stew- art, Larry Angel, Mike Finnegan, Harriet Ottenheimer, Marty Ottenheimer, Roberto Frisancho, Randy Susman, Karen Strier, Joanna Lambert, Jim Hijiya, Cecil Brown, Bill Fash, Rich Blanton, Henry Wright, James Griffin, Bill Jungers, David Frayer, Bill Pollitzer, George Armelagos, Jane Buikstra, Elwyn Simons, Steve Churchill, Neil Tubbs, Bob Bettinger, Tim White, Dean Falk, Owen Lovejoy, Scott Simpson, David Carlson, Alan Goodman, Bill Dancey, Debbie Guatelli- Steinberg, Sam Stout, Clark Mallam, and Chris Peebles.

I would like to thank Joanna E. Lambert, University of Colorado–Boulder and Friderun Ankel-Simons, Duke Uni- versity for their help and their words used to prepare the back cover description. Their response was helpful, timely, and their suggested wording was perfect.

The book benefited from the expertise of many anthro- pologists and other experts. I especially acknowledge the fol- lowing reviewers for their insights, advice, and suggestions for revision of the text and creation of the support package:

Sabrina Agarwal, University of California, Berkeley Paul Aiello, Ventura College Lon Alterman, North Carolina State University

Tara Devi Ashok, University of Massachusetts Boston Diana Ayers- Darling, Mohawk Valley Community

College Philip de Barros, Palomar College Thad Bartlett, University of Texas at San Antonio Cynthia Beall, Case Western Reserve University Owen Beattie, University of Alberta Anna Bellisari, Wright State University Daniel Benyshek, University of Nevada, Las Vegas Tracy Betsinger, State University of New York at Oneonta Deborah Blom, University of Vermont Amy Bogaard, Oxford University Günter Bräuer, University of Hamburg Emily Brunson, University of Washington Victoria Buresch, Glendale Community College Isabelle Champlin, University of Pittsburgh at Bradford Joyce Chan, California State University, Dominguez

Hills Chi- hua Chiu, Kent State University David Clark, Catholic University of America Robert Corruccini, Southern Illinois University Herbert Covert, University of Colorado Douglas Crews, Ohio State University Eric Delson, Lehman College, City University of

New York Katherine Dettwyler, University of Delaware Joanne Devlin, University of Tennessee Paul Erickson, St. Mary’s University Becky Floyd, Cypress College David Frayer, University of Kansas Daniel Gebo, Northern Illinois University Anne Grauer, Loyola University of Chicago Mark Griffin, San Francisco State University Michael Grimes, Western Washington University Gregg Gunnell, Duke University Lesley Harrington, University of Alberta Lauren Hasten, Las Positas College John Hawks, University of Wisconsin– Madison Samantha Hens, California State University, Sacramento James Higham, New York University Madeleine Hinkes, San Diego Mesa College Homes Hogue, Ball State University Nina Jablonski, Pennsylvania State University Karin Enstam Jaffe, Sonoma State University Gabriela Jakubowska, Ohio State University Gail Kennedy, University of California, Los Angeles Dawn Kitchen, Ohio State University Haagen Klaus, George Mason University Patricia Lambert, Utah State University Michael Little, Binghamton University Chris Loeffler, Irvine Valley College

xxvi Instructor

 

 

Sara Lynch, Queens College, City University of New York Lorena Madrigal, University of South Florida Ann Magennis, Colorado State University Stephen Marshak, University of Illinois at

Urbana– Champaign Debra Martin, University of Nevada, Las Vegas Thomas McDade, Northwestern University William McFarlane, Johnson County Community

College Scott McGraw, Ohio State University Rachel Messinger, Moorpark College Ellen Miller, Wake Forest University Leonor Monreal, Fullerton College Ellen Mosley- Thompson, Ohio State University Michael Muehlenbein, Indiana University Dawn Neill, California Polytechnic State University, San

Luis Obispo Wesley Niewoehner, California State University, San

Bernardino Kevin Nolan, Ball State University Rachel Nuger, Hunter College, City University of

New York Dennis O’Rourke, University of Utah Janet Padiak, McMaster University Amanda Wolcott Paskey, Cosumnes River College Michael Pilakowski, Butte College Janine Pliska, Golden West College Deborah Poole, Austin Community College Leila Porter, Northern Illinois University Frances E. Purifoy, University of Louisville Mary Ann Raghanti, Kent State University Lesley M.  Rankin- Hill, University of Oklahoma Jeffrey Ratcliffe, Penn State Abington Laurie Reitsema, University of Georgia Melissa Remis, Purdue University Analiese Richard, University of the Pacific Charles Roseman, University of Illinois Karen Rosenberg, University of Delaware John Rush, Sierra College Andrew Scherer, Brown University Timothy Sefczek, Ohio State University Lynette Leidy Sievert, University of Massachusetts

Scott W. Simpson, Case Western Reserve University Cynthia Smith, Ohio State University Fred Smith, Illinois State University Richard Smith, Washington University Sara Smith, Delta College Lilian Spencer, Glendale Community College Sara Stinson, Queens College, City University of

New York Christopher Stojanowski, Arizona State University Margaret Streeter, Boise State University Karen Strier, University of Wisconsin– Madison Nancy Tatarek, Ohio University Lonnie Thompson, Ohio State University Victor Thompson, University of Georgia Christopher Tillquist, University of Louisville Sebina Trumble, Hartnell College Lisa Valkenier, Merritt College Dennis Van Gerven, University of Colorado Boulder Ronald Wallace, University of Central Florida David Webb, Kutztown University Daniel Wescott, Texas State University Tim White, University of California, Berkeley Janet Wiebold, Spokane Community College Caleb Wild, Mira Costa College Leslie Williams, Beloit College Sharon Williams, Purdue University Kristen Wilson, Cabrillo College Milford Wolpoff, University of Michigan Thomas Wynn, University of Colorado Colorado Springs

Thanks, everyone, for your help! Lastly, a very special thanks goes to all of the faculty around the globe who adopted the previous two editions of Essentials of Physical Anthropology for their introductory physical anthropology classes. I am also grateful to the hundreds of students who connected with the book— many of whom have written me with their comments. Please continue to send me your com- ments ([email protected]).

Columbus, Ohio August 10, 2015

Instructor xxvii

 

 

xxviii

TO THE STUDENT

PHYSICAL ANTHROPOLOGY IS ABOUT DISCOVERING WHO WE ARE

THINKING LIKE AN ANTHROPOLOGIST

Who are we? Where do we come from? Why do we look and act the way we do? This book is a journey that addresses these and other big questions about us, Homo sapiens. This journey emphasizes humans’ discovery of the fascinating record of our diversity and of our evolution, a record that serves as a collective memory of our shared biological pres- ence on Earth. From here to the end of the book, I will share with you all kinds of ideas that add up to our current understanding of human beings as living organisms. Along the way, you will experience scientific breakthroughs such as the Human Genome Project and forensics (you might even watch CSI and Bones in a whole new way). You will gain new understandings of phenomena such as race and human diversity, global warming and its impact on our evolution and our well- being, the origins of human violence, global disease, and the growing worldwide obesity epidemic. Like an anthropologist tackling important questions, you will discover places on nearly every continent and come to see what life was like for millions of years before the present, before the emergence and evolution of humans.

Neither your instructor nor I can expect you as an intro- ductory student to understand all the developments in phys- ical anthropology. Both of us can, however, present you with a clear and concise framework of the field. By the time you are finished reading this book and completing this course, you will have a solid background in the basic tenets of the discipline. This knowledge will help you understand your place in nature and the world that we— more than 7 billion of us and growing— live in. The framework for developing your understanding of physical anthropology is  the scien- tific method, a universal approach to understanding the very complex natural world. You should not assume that this book and this course are about only knowing the right

answers, the “facts” of physical anthropology. Rather, they are also about seeing how physical anthropologists know what they know— understanding the scientific method. So as you read, keep in mind the key questions that scientists try to answer, their processes and methods for finding the answers, and the answers themselves.

In writing this book, I have focused on the big ques- tions in physical anthropology, how scientists have tackled them, and what key discoveries have been made. I have not shied away from identifying the scientists who made these discoveries— real people, young and old, from all over the world. Whether you need to learn all these individuals’ names and what they contributed to the growth of physical anthropology and to our knowledge of human evolution and variation is up to your instructor. But in the introductory physical anthropology class that I teach, I encourage my stu- dents to learn about the people behind the ideas. By seeing the field through these people’s eyes, you can start thinking like an anthropologist.

SEEING LIKE AN ANTHROPOLOGIST

Thinking like an anthropologist includes seeing what anthropologists see. We anthropologists are constantly looking at things— fossilized human teeth, ancient DNA, excavated stone tools, primate skeletons, and much more— and using what we see to understand biology in the past and in the present. The photos and drawn art throughout this book have been chosen to help you see what anthropologists see. I strongly encourage you to pay close attention to the visuals in the book and their captions because much of our anthropological understanding is in the art program.

THE STRUCTURE OF THE BOOK

The book is divided into two parts. Following an overview of anthropology and physical anthropology (chapter 1), Part I provides the basic context for how we understand human (and our nonhuman primate relatives’) biology in the present

 

 

(and how that helps us understand the past). From this section of the book you should come away with an under- standing of evolution and the biology associated with it. Evolution as an idea has a long history (chapter 2). You will need to fully grasp the meaning and power of this theory, which explains humans’ biological variation today and in the past. Part I also has the important job of providing you with an understanding of genetics (chapters  3 and 4). This information is a central part of the evidence for evolution, from the level of the molecule to the level of the population.

Part I also looks at the biology of living people, that of the other living primates, and the variation among primate spe- cies. I am keen on debunking the common notion that there are discrete categories— races— of human beings (chapter 5). In fact, nothing about the biology of people, present or past, indicates that we can be divided into distinct groups. After looking at how environment and culture help shape the way humans look and behave, I will look similarly at nonhuman primates (chapters  6 and 7). Because nonhuman primates’ appearances are much more categorical than humans’ are, nonhuman primate appearance lends itself to classification or taxonomy. In these chapters, we will look at what nonhu- man primates do in the wild, what they are adapted to, and especially the environment’s role in shaping their behavior and biology. By looking at living people and living nonhu- man primates, we will be better equipped to understand the biological evidence drawn from the past.

Part II examines the processes and evidence physical anthropologists and other scientists use to understand the past (chapter  8), the evolution of prehuman primate ances- tors that lived more than 50  million years ago (chapter  9), and both the emergence of our humanlike ancestors and their evolution into modern humans (chapters  10, 11, and 12). Contrary to popular (and some scientific) opinion, human evolution did not stop when anatomically modern people first made their appearance in various corners of the globe. Rather, even into the last 10,000 years a considerable amount of biological change has occurred. Anthropologists have learned that agriculture, which began some 10,000 years ago, has been a fundamental force behind population increase. The downside of this shift to new kinds of food and the resulting population increase was a general decline in health. The later section of Part II (chapter 13) explores the nature and cause of biological change, including the changes associated with health and well- being that led to the biological and environmental conditions we face today.

With this book in hand and our goals— thinking and see- ing like anthropologists— in mind, let us set off on this excit- ing journey. Consider it a voyage of discovery, on which our shipmates include your instructor and your fellow students. If we work hard and work together, we will find perhaps the most interesting thing on Earth: ourselves.

To the Student xxix

Gorilla meets hominin and author of Essentials of Physical Anthropology Clark Larsen.

 

 

 

ESSENTIALS OF PHYSICAL ANTHROPOLOGY

 

 

THE GEORGIA COAST WAS A FOCAL point for Spanish colonization in the sixteenth and seventeenth centuries. European colo- nization set in motion changes in human living conditions that eventually affected human biology on a global scale.

 

 

3 3

1 What is anthropology?

What is physical anthropology?

What makes us human and different from other animals?

How do physical anthropologists know what they know?

1

2

3

4

What Is Physical Anthropology?

In the heat of the midday summer sun, our boat slowly made its way across the five miles of water that separate mainland Georgia from St. Catherines Island, one of a series of barrier islands dotting the Atlantic seaboard. Today, the island is covered by dense vegetation typical of the subtropical American South— palmettos and other palm trees, pines, hickories, and live oaks— and is infested with a wide array of stinging and biting insects. It is hard to imagine that this setting was once a focal point of the Span- ish colonial “New World,” representing the northernmost extension of Spain’s claim on eastern North America. This was the location of the Roman Catholic church and mis- sion Santa Catalina de Guale, where several hundred Indians and a dozen Spaniards lived and worked during the late 1500s and most of the 1600s.

What could possibly have motivated my field team and me to work for months under a blazing sun, fighting insects? Like any scientific investigation, our fieldwork was moti- vated by specific questions that we keenly wanted to answer. Buried in the sands of St. Catherines were the mortal remains— skeletons— of the native people who had lived at this long- abandoned place. These remains held answers to questions about the biology of modern people. Native Americans had lived in this area of the world for most of the last 10,000 years. We wanted to know about their biological evolution and vari- ation: How had these people changed biologically over this time span? What caused these changes? What circumstances led to the changes that we hoped to identify and interpret?

When we first set foot on St. Catherines Island in the summer of 1982 to begin our work at Mission Santa Catalina, we were excited about our project, but little did we

B I G Q U E S T I O N S ?

 

 

4 | CHAPTER 1 What Is Physical Anthropology?

realize just what a spectacular scientific journey we were undertaking. The skel- etons we sought turned out to provide wonderfully rich biological details about a little- understood region of the world, especially relating to the health consequences and behavioral consequences of European contact on native peoples. In setting up the research project, I had envisioned that our findings would provide a microcosm of what had unfolded globally— in the Americas, Asia, Africa, and Australia— during the previous 500 years of human history. During this period, significant biological changes had taken place in humans. Some of these changes were evolutionary— they resulted in genetic change. Other biological changes, nonevolutionary ones, reflected significant alterations in health and lifestyle, alterations that had left impressions on the skeletons we studied. Such study— of genetic and nongenetic changes— here and elsewhere in the world has proven fundamental to human beings’ understanding of their biology in the early twenty- first century.

Like any scientific investigation, the research project at Mission Santa Catalina did not develop in a vacuum. Prior to our work there, my team and I had devoted nearly a decade to studying hundreds of skeletons we had excavated from the region, dating from before the arrival of Spaniards. We had learned from archaeological evidence that before AD 1000 or so the people there ate exclusively wild animals, fish, and wild plants— they were hunters and gatherers. Never settling into one place for any period of time, they moved from place to place over the year, hunting animals, fishing on the coastline, and collecting plants. Then, their descendants— the ancestors of the mission Indians— acquired corn agriculture, becoming the first farmers in the region. These people did lots of fishing, but farming produced the mainstay of their diet. This major shift in lifestyle led to the establishment of semipermanent villages. In comparison with the hunter– gatherers living before AD 1000, the later agricultural people were shorter, their skulls and limb bones were smaller, and they had more dental disease and more infections. All of this information— scientific discoveries about the prehistoric people, their biological changes, and their adaptations— set the stage for our return to the island to study the people who lived at Santa Catalina, the descendants of the prehistoric hunter– gatherers and later farmers. From our study of their remains, we learned that after the Spaniards’ arrival the native people worked hard, they became more focused on producing and eating corn, and their health declined. The combination of declining quality of life and new diseases introduced by the Spaniards led to the native people’s extinction in this area of North America.

The research just described is one small part of the broader discipline known as physical anthropology. My work concerns life on the southeastern United States Atlantic coast, but physical anthropologists explore and study everywhere how humans and their ancestors lived. This enterprise covers a lot of ground and a lot of time, basi- cally the entire world and the last 50 million years or so! The territorial coverage of physical anthropology is so widespread and so diverse because the field addresses broad issues, seeking to understand human evolution— what we were in the past, who we are today, and where we will go in the future. Physical anthropologists seek answers to questions about why we are what we are as biological organisms. How we answer these questions is oftentimes difficult. The questions, though, motivate physical anthropologists to spend months in the subtropics of coastal Georgia, learning about an extinct native people; in the deserts of central Ethiopia, finding and studying the remains of people who lived hundreds, thousands, or even millions

 

 

What Is Anthropology? | 5

of years ago; or at the high altitudes of the Andes Mountains, studying living people and their responses and long- term adaptation to low oxygen and extreme cold, to name just a few of the settings you will learn about in this book. In this chapter, we will explore in more detail the nature of physical anthropology and its subject matter.

What Is Anthropology? When European explorers first undertook transcontinental travel (for example, Marco Polo into Asia in the late 1200s) or transoceanic voyages to faraway lands (for example, Christopher Columbus to the Americas in the late 1400s and early 1500s), they encountered people that looked, talked, dressed, and behaved very differently from themselves. When these travelers returned to their home coun- tries, they described the peoples and cultures they saw. Building on these accounts, early scholars speculated on the relationships between humans living in Europe and those encountered in distant places. Eventually, later scholars developed new ideas about other cultures, resulting in the development of the discipline of anthropology.

Anthropology is the study of humankind, viewed from the perspective of all people and all times. As it is practiced in the United States, it includes four branches or subdisciplines: cultural anthropology, archaeology, linguistic anthropology, and physical anthropology, also called biological anthropology (Figure 1.1).

Cultural anthropologists typically study present- day societies in non- Western settings, such as in Africa, South America, or Australia. Culture— defined as learned behavior that is transmitted from person to person— is the unifying theme of study in cultural anthropology.

Archaeologists study past human societies, focusing mostly on their material remains— such as animal and plant remains and places where people lived in the past. Archaeologists are best known for their study of material objects— artifacts— from past cultures, such as weaponry and ceramics. Archaeologists study the processes behind past human behaviors— for example, why people lived where they did, why some societies were simple and others complex, and why people shifted from hunting and gathering to agriculture beginning more than 10,000 years ago. Archaeologists are the cultural anthropologists of the past— they seek to reassemble cultures of the past as though those cultures were alive today.

Linguistic anthropologists study the construction and use of language by human  societies. Language— defined as a set of written or spoken symbols that refer to things (people, places, concepts, etc.) other than themselves— makes possible the transfer of knowledge from one person to the next and from one gen- eration to the next. Popular among linguistic anthropologists is a subfield called sociolinguistics, the investigation of language’s social contexts.

Physical (or biological) anthropologists study all aspects of present and past human biology. As we will explore in the next section, physical anthropology deals with the evolution of and variation among human beings and their living and past relatives.

No anthropologist is expected to be an expert in all four branches. Anthropol- ogists in all four areas and with very different interests, however, acknowledge the diversity of humankind in all contexts. No other discipline embraces the breadth of the human condition in this manner. In fact, this remarkably diverse discipline

anthropology The study of humankind, viewed from the perspectives of all peo- ple and all times.

cultural anthropology The study of modern human societies through the analysis of the origins, evolution, and variation of culture.

archaeology The study of historic of pre- historic human populations through the analysis of material remains.

linguistic anthropology The study of the construction, use, and form of language in human populations.

physical anthropology The original term for biological anthropology.

biological (physical) anthropology The study of the evolution, variation, and adaptation of humans and their past and present relatives.

culture Learned behavior that is transmit- ted from person to person.

artifacts Material objects from past cultures.

language A set of written or spoken sym- bols that refer to things (people, places, concepts, etc.) other than themselves.

sociolinguistics The science of investigat- ing language’s social contexts.

 

 

6 | CHAPTER 1 What Is Physical Anthropology?

differs from other disciplines in its commitment to the notion that, unlike other animals, humans are biocultural— both biological and cultural beings. Anthro- pologists are interested in the interrelationship between biology and culture. Anthropologists call this focus the biocultural approach. Anthropology also differs from other disciplines in emphasizing a broad comparative approach to the study of biology and culture, looking at all people (and their ancestors) and all cultures in all times and all places. Anthropologists are interested in people and their ancestors, wherever or whenever they lived. Simply, you are studying a field that is holistic, unlike any you have studied before.

biocultural approach The scientific study of the interrelationship between what humans have inherited genetically and culture.

The Four Branches of Anthropology

Cultural Anthropology Archaeology Linguistic Anthropology Physical Anthropology

The study of cultures and societies of human beings and their very recent past.

Traditional cultural anthropologists study

living cultures and present their observations in

an ethnography.

The study of past societies and their cultures,

especially the material remains of the past, such

as tools, food remains, and places where people lived.

The study of language, especially how language

is structured, the evolution of language, and the social

and cultural contexts for language.

Also called biological anthropology, physical

anthropology is the study of human evolution and variation, both past and current.

FIGURE 1.1 The Four Branches of Anthropology (a) Cultural anthropologists, who study living populations, often spend time living with cultural groups to gain more intimate perspectives on those cultures. The American anthropologist Margaret Mead (1901–1978), one of the most recognizable names in cultural anthropology, studied the peoples of the Admiralty Islands, near Papua New Guinea. (b) Archaeologists study past human behaviors by investigating material remains that humans leave behind, such as buildings and other structures. In the Peruvian Andes, this archaeologist examines the remnants of a brewery used by the Wari Empire (ca. AD 750–1000). (c) Linguistic anthropologists study all aspects of language and language use. Here, Leslie Moore, a linguistic anthropologist working in a Fulbe community in northern Cameroon, records as a teacher guides a boy in memorizing Koranic verses. (d) Physical anthropologists study human evolution and variation. Some physical anthropologists study skeletons from the past to investigate evolution and variation throughout human history. Those working in forensic anthropology, a specialty within physical anthropology, examine skeletons to identify who they were in life. Such an identification may be of a single person or of thousands. For example, the forensic anthropologist pictured here was called on to help identify the estimated 30,000 victims of Argentina’s “Dirty War,” which followed the country’s 1976 coup.

(a) (b) (c) (d)

forensic anthropology The scientific exam- ination of skeletons in hope of identifying the people whose bodies they came from.

 

 

What Is Physical Anthropology? | 7

What Is Physical Anthropology? The short answer to this question is, Physical anthropology is the study of human biologi- cal evolution and human biocultural variation. Two key concepts underlie this definition.

Number one, every person is a product of evolutionary history, or all the bio- logical changes that have brought humanity to its present form. The remains of humanlike beings, or hominins, indicate that the earliest human ancestors, in Africa, date to sometime around 6–8  million years ago (mya). Since that time, the physical appearance of hominins and their descendants, including modern humans, has changed dramatically. Our physical appearance, our intelligence, and everything else that makes us distinctive biological organisms evolved in our predecessors, whose genes led to the species we are today. (Genes and species are among the subjects of chapters 3 and 4.)

Number two, each of us is the product of his or her own individual life history. From the moment you were conceived, your biological makeup has been deter- mined mostly by your genes. (The human genome— that is, all the genetic mate- rial in a person— includes some 20,000–25,000 genes.) Your biological makeup is also strongly influenced by your environment. Environment here refers not just to the obvious factors such as climate but to everything that has affected you— the physical activities you have engaged in (which have placed stress on your muscles and bones), the food you have eaten, and many other factors that affect overall health and well- being. Environment also includes social and cultural factors. A disadvantaged social environment, such as one in which infants and children receive poor- quality nutrition, can result in negative consequences such as poor health, reduced height, and shortened life expectancy. The Indian child who lived after the shift from foraging to farming on the Georgia coast ate more corn than did the Indian child who lived in the same place before AD 1000. Because of the corn- rich diet, the later child’s teeth had more cavities. Each child’s condition reflects millions of years of evolution as well as more immediate circumstances, such as diet, exposure to disease, and the stresses of day- to- day living.

WHAT DO PHYSICAL ANTHROPOLOGISTS DO? Physical anthropologists routinely travel to places throughout the United States and around the world to investigate populations. Some physical anthropologists study living people, while others study extinct and living species of our nearest biological relatives, primates such as lemurs, monkeys, and apes. I am among the physical anthropologists who travel to museum collections and archaeological localities to study past societies. When I tell people outside the field what I do for a living, they often think physical anthropology is quite odd, bizarre even. Frequently they ask, “Why would anyone want to study dead people and old bones and teeth?” Everyone has heard of physics, chemistry, and biology; but the average person has never heard of this field. Compared to other areas of science, physical anthropology is small. But smallness does not make it unimportant. It is practical and important, providing answers to fundamental questions that have been asked by scholars and scientists for centuries, such as Who are we as a species? What does it mean to be human? Where did we come from? Moreover, physical anthropology plays a vital role in address- ing questions that are central to our society, sometimes involving circumstances that all of us wish had never come about. For example, the tragedy that Americans identify as 9/11 called immediately for the assistance of specialists from forensic anthropology.

hominin Humans and humanlike ancestors.

genome The complete set of genetic information— chromosomal and mito- chondrial DNA— for an organism or spe- cies that represents all of the inheritable traits.

primates A group of mammals in the order Primates that have complex behavior, varied forms of locomotion, and a unique suite of traits, including large brains, forward- facing eyes, fingernails, and reduced snouts.

 

 

8 | CHAPTER 1 What Is Physical Anthropology?

The discipline as practiced in the United States began in the first half of the twentieth century, especially under the guidance of three key figures: Franz Boas for American anthropology generally; Czech- born Aleš HrdliČka, who started the professional scientific journal and professional society devoted to the field; and Earnest Hooton, who trained most of the first generation of physical anthro- pologists. While the theory and methods of physical anthropologists today have changed greatly since the early 1900s, the same basic topics first envisioned by these founders form what we do.

Physical anthropologists study all aspects of human biology, specifically looking at the evolution and variation of human beings and their living and past relatives. This focus on biology means that physical anthropologists practice a biological science. But they also practice a social science, in that they study biology within the context of culture and behavior. Depending on their areas of interest, physical anthropologists might examine molecular structure, bones and teeth, blood types, breathing capacity and lung volume, genetics and genetic history, infectious and other types of disease, origins of language and speech, nutrition, reproduction, growth and development, aging, primate origins, primate social behavior, brain biology, and many other topics dealing with variation in both the living and the dead— sometimes the very long dead (Figure 1.2)!

In dealing with such topics, physical anthropologists apply methods and theo- ries developed in other disciplines as well as in their own as they answer questions that help us understand who we are, a point that I will raise over and over again throughout this book. The very nature of their discipline and their constant borrowing from other disciplines mean that physical anthropologists practice an interdisciplinary science. For example, they might draw on the work of geologists who study the landforms and layering of deposits of soil and rock that tell us when earlier humans lived. Or they might obtain information from paleontologists, who study the evolution of life- forms in the distant past and thus provide the essential context for understanding the world in which earlier humans lived. Some physical anthropologists are trained in chemistry, so they can analyze the chemical properties of bones and teeth to determine what kinds of foods were eaten by those earlier humans. Or to learn how living humans adapt to reduced- oxygen settings, such as in the high altitudes of the Peruvian Andes Mountains, physical anthropologists might work with physiologists who study the lungs’ ability to absorb oxygen. The firm yet flexible identity of their science allows physical anthropologists to gather data from other disciplines in order to address key questions. Questions drive what they do.

What Makes Humans So Different from Other Animals?: The Six Steps to Humanness Human beings clearly differ from other animals. From humanity’s earliest origin— about 6–8 mya, when an apelike primate began walking on two feet— to the period beginning about 10,000 years ago, when modern climates and environments emerged following what is commonly known as the Ice Age, six key attributes developed that make us unique. These attributes are bipedalism, nonhoning chewing, complex material culture and tool use, hunting, speech, and dependence on domesticated foods (Figure 1.3, pp. 10–11). The first development represents

FIGURE 1.2 A Sample of What Physical Anthropologists Do (a) Human remains excavated at Bactia Pozzeveri, a medieval church cemetery in Tuscany, Italy, provide a window onto health and living conditions in Europe. (b) Geneticists analyze samples of human DNA for various anthropological purposes. DNA studies are used to determine how closely related humans are to other primate species, to examine human origins, and to determine individual identities. (c) A human biologist records the physical activities of a lactating woman (right, weaving basket) living in a rural community in the eastern Amazon, Brazil. These data will be used to calculate the woman’s energy expenditure and to understand how she copes with reproduction’s great energy demands. (d) In a lab, a forensic anthropologist measures and assesses human bones. If the bones came from a contemporary grave, this forensic information might help to identify the victim. If the bones belonged to a past population, physical anthropologists might use these data to gain insight into the population’s health and lifestyle. (e) Laboratory investigations of human ancestors’ bones help paleoanthropologists to determine where these ancestors fit in the human family tree. (f) Primatologists, such as the British researcher Jane Goodall (b. 1934), study our closest living relatives, nonhuman primates. The behavior and lifestyle of chimpanzees, for example, help physical anthropologists to understand our evolutionary past.

(c)

(a)

(e)

(b)

(d)

(f)

 

 

What Makes Humans So Different from Other Animals?: The Six Steps to Humanness | 9

The discipline as practiced in the United States began in the first half of the twentieth century, especially under the guidance of three key figures: Franz Boas for American anthropology generally; Czech- born Aleš HrdliČka, who started the professional scientific journal and professional society devoted to the field; and Earnest Hooton, who trained most of the first generation of physical anthro- pologists. While the theory and methods of physical anthropologists today have changed greatly since the early 1900s, the same basic topics first envisioned by these founders form what we do.

Physical anthropologists study all aspects of human biology, specifically looking at the evolution and variation of human beings and their living and past relatives. This focus on biology means that physical anthropologists practice a biological science. But they also practice a social science, in that they study biology within the context of culture and behavior. Depending on their areas of interest, physical anthropologists might examine molecular structure, bones and teeth, blood types, breathing capacity and lung volume, genetics and genetic history, infectious and other types of disease, origins of language and speech, nutrition, reproduction, growth and development, aging, primate origins, primate social behavior, brain biology, and many other topics dealing with variation in both the living and the dead— sometimes the very long dead (Figure 1.2)!

In dealing with such topics, physical anthropologists apply methods and theo- ries developed in other disciplines as well as in their own as they answer questions that help us understand who we are, a point that I will raise over and over again throughout this book. The very nature of their discipline and their constant borrowing from other disciplines mean that physical anthropologists practice an interdisciplinary science. For example, they might draw on the work of geologists who study the landforms and layering of deposits of soil and rock that tell us when earlier humans lived. Or they might obtain information from paleontologists, who study the evolution of life- forms in the distant past and thus provide the essential context for understanding the world in which earlier humans lived. Some physical anthropologists are trained in chemistry, so they can analyze the chemical properties of bones and teeth to determine what kinds of foods were eaten by those earlier humans. Or to learn how living humans adapt to reduced- oxygen settings, such as in the high altitudes of the Peruvian Andes Mountains, physical anthropologists might work with physiologists who study the lungs’ ability to absorb oxygen. The firm yet flexible identity of their science allows physical anthropologists to gather data from other disciplines in order to address key questions. Questions drive what they do.

What Makes Humans So Different from Other Animals?: The Six Steps to Humanness Human beings clearly differ from other animals. From humanity’s earliest origin— about 6–8 mya, when an apelike primate began walking on two feet— to the period beginning about 10,000 years ago, when modern climates and environments emerged following what is commonly known as the Ice Age, six key attributes developed that make us unique. These attributes are bipedalism, nonhoning chewing, complex material culture and tool use, hunting, speech, and dependence on domesticated foods (Figure 1.3, pp. 10–11). The first development represents

FIGURE 1.2 A Sample of What Physical Anthropologists Do (a) Human remains excavated at Bactia Pozzeveri, a medieval church cemetery in Tuscany, Italy, provide a window onto health and living conditions in Europe. (b) Geneticists analyze samples of human DNA for various anthropological purposes. DNA studies are used to determine how closely related humans are to other primate species, to examine human origins, and to determine individual identities. (c) A human biologist records the physical activities of a lactating woman (right, weaving basket) living in a rural community in the eastern Amazon, Brazil. These data will be used to calculate the woman’s energy expenditure and to understand how she copes with reproduction’s great energy demands. (d) In a lab, a forensic anthropologist measures and assesses human bones. If the bones came from a contemporary grave, this forensic information might help to identify the victim. If the bones belonged to a past population, physical anthropologists might use these data to gain insight into the population’s health and lifestyle. (e) Laboratory investigations of human ancestors’ bones help paleoanthropologists to determine where these ancestors fit in the human family tree. (f) Primatologists, such as the British researcher Jane Goodall (b. 1934), study our closest living relatives, nonhuman primates. The behavior and lifestyle of chimpanzees, for example, help physical anthropologists to understand our evolutionary past.

(c)

(a)

(e)

(b)

(d)

(f)

 

 

BIPEDALISM 6 MYA

The upright, bipedal (two-footed) gait was the first hallmark feature of our hominin ancestors.

SPEECH 2.5 MYA

In the entire animal kingdom, only humans can speak and, through speech, express complex thoughts and ideas. The shape of the hyoid bone is unique to hominins and reflects their ability to speak. Speech is part of the overall package in the human lineage of increased cognition, intelligence, and brain-size expansion.

HUNTING 1 MYA

Humans’ relatively large brains require lots of energy to develop and function. Animal protein is an ideal source of that energy, and humans obtained it for most of their evolution by eating animals they hunted. To increase their chances of success in hunting, humans employed tools they made and cooperative strategies.

NONHONING CHEWING 5.5 MYA

Humans’ nonhoning chewing complex (below) lacks large, projecting canines in the upper jaw and a diastema, or gap, between the lower canine and the third premolar.

The chewing complex of apes such as gorillas (below) has large, projecting upper canines and a diastema in the lower jaw to accommodate them.

MATERIAL CULTURE AND TOOLS 3.3 MYA

Humans’ production and use of stone tools is one example of complex material culture. The tools of our closest living relatives, the chimpanzees, do not approach the complexity and diversity of modern and ancestral humans’ tools

DOMESTICATED FOODS 11,000 YEARS AGO

In recent evolution—within the last 10,000 years or so—humans domesticated a wide variety of plants and animals, controlling their life cycles and using them for food and other products, such as clothing and shelter.

Human Ape Diastema

Hyoid bone

F I G U R E

1.3 The Six Big Events of Human Evolution: Bipedalism, Nonhoning Chewing, Dependence on Material Culture, Speech, Hunting, and Domestication of Plants and Animals

 

 

BIPEDALISM 6 MYA

The upright, bipedal (two-footed) gait was the first hallmark feature of our hominin ancestors.

SPEECH 2.5 MYA

In the entire animal kingdom, only humans can speak and, through speech, express complex thoughts and ideas. The shape of the hyoid bone is unique to hominins and reflects their ability to speak. Speech is part of the overall package in the human lineage of increased cognition, intelligence, and brain-size expansion.

HUNTING 1 MYA

Humans’ relatively large brains require lots of energy to develop and function. Animal protein is an ideal source of that energy, and humans obtained it for most of their evolution by eating animals they hunted. To increase their chances of success in hunting, humans employed tools they made and cooperative strategies.

NONHONING CHEWING 5.5 MYA

Humans’ nonhoning chewing complex (below) lacks large, projecting canines in the upper jaw and a diastema, or gap, between the lower canine and the third premolar.

The chewing complex of apes such as gorillas (below) has large, projecting upper canines and a diastema in the lower jaw to accommodate them.

MATERIAL CULTURE AND TOOLS 3.3 MYA

Humans’ production and use of stone tools is one example of complex material culture. The tools of our closest living relatives, the chimpanzees, do not approach the complexity and diversity of modern and ancestral humans’ tools

DOMESTICATED FOODS 11,000 YEARS AGO

In recent evolution—within the last 10,000 years or so—humans domesticated a wide variety of plants and animals, controlling their life cycles and using them for food and other products, such as clothing and shelter.

Human Ape Diastema

Hyoid bone

 

 

12 | CHAPTER 1 What Is Physical Anthropology?

the most profound physical difference between humans and other animals, namely the manner in which we get around: we are committed to bipedalism, that is, walking on two feet. The next development was the loss of a large, honing canine tooth, like the one that apes typically use to shred their food (mostly plants), to the simple nonhoning canine, with which we simply process food. Our ancestors’ honing canine disappeared because they developed the ability to make and use tools for processing food.

Today, our species completely depends on culture— and especially material culture— for its day- to- day living and its very survival. Culture is a complex human characteristic that facilitates our survival by enabling us to adapt to different settings. Material culture is the part of culture that is expressed as objects that humans use to manipulate environments. For example, hammers and nails are forms of material culture that enable us to make cabinets, tables, and countless other forms of material culture. The material remains of past cultures go back hundreds of thousands of years, to the first simple tools made from rocks 3.3 mya (Figure 1.4). Material culture today makes our lifestyles possible. Can you imag- ine your life without it? We could survive without modern additions to material culture, such as cars, computers, TVs, plumbing, and electricity, as our ancestors did before the last century. What about living without basic material culture, such as shelter and clothing, especially in climates where it can be very, very cold in the winter? Without material culture, how would any of us get food? The answer to both questions is simple: we could not make it without some forms of technology— to regulate temperature, to acquire food, and so on. Some societies are much less technologically complex than others, but no society functions with- out any technology.

Anthropologists and animal behaviorists have shown that human beings are not, however, the only type of animal that has or can employ material culture. Primatologists have observed some chimpanzee societies in Africa, for example, making simple tools from twigs (Figure  1.5). In laboratories, chimpanzees have been taught to use physical symbols that approximate human communication. Still, these and other forms of material culture used by nonhuman species are nowhere near as complex as those created by humans.

The other three key attributes of humanness— hunting, speech, and depen- dence on domesticated foods— appeared much later in human evolution than bipedalism, nonhoning chewing, and complex material culture and tool use. Hunting here refers to the social behavior whereby a group, adult men in general, organize themselves to pursue animals for food. This behavior likely dates back to a million or more years ago. Some nonhuman primates organize to pursue prey, but they do not use tools or travel long distances as humans distinctively do when they hunt.

An equally distinctive human behavior is speech. We are the only animal that communicates by talking. Unfortunately for research purposes, recording- and- listening technology was invented only about a century ago. For information about long- past speech, anthropologists rely on indirect evidence within the skeleton. For example, the hyoid bone, in the neck, is part of the vocal structure that helps produce words. The unique appearance of the human hyoid helps anthropologists conjecture about the origins of speech.

The most recently developed unique human behavior is the domesticated manner in which we acquire our food. About 10,000–11,000 years before the pres- ent (yBP), humans began to raise animals and grow plants. This development led

bipedalism Walking on two feet.

nonhoning canine An upper canine that, as part of a nonhoning chewing mechanism, is not sharpened against the lower third premolar.

material culture The part of culture that is expressed as objects that humans use to manipulate environments.

FIGURE 1.4 First Tools The earliest stone tools date to 3.3 mya and are associated with early human ancestors in East Africa. The example shown here is from Lomekwi, West Turkana, Kenya. This tool had various functions, including the processing of plants and meat for food.

 

 

What Makes Humans So Different from Other Animals?: The Six Steps to Humanness | 13

to our current reliance on domesticated species. This reliance has had a profound impact on human biology and behavior and represents a pivotal step in human evolution.

Human beings’ unique behaviors and survival mechanisms, and the anatomical features related to them, arose through the complex interaction of biology and culture. Indeed, our ancestors’ increasing dependence on culture for survival has made us entirely culture- dependent for survival. The behaviors that are unique to humans— speech, tool use, and dependence on culture— are also related to the fact that humans are very smart. Our remarkable intelligence is reflected in our abilities to think and interact in the ways we do (and take for granted), to communicate in complex ways, and to accomplish diverse tasks on a daily basis to survive. Our brains are bigger and have more complex analytical skills than do the brains of both other primates and animals in general. These biological advantages enable us to figure out complex problems, including how to survive in a wide range of environments.

The American anthropologist Robert Boyd and his colleagues argue that while humans are the smartest animals, in no way are we individually smart enough to acquire all the complex information necessary to survive in any particular envi- ronment. Today and through much of human evolution, our species has survived owing to our complex culture, including tool use and technology generally, prac- tices, and beliefs. For hundreds of thousands of years, humans have had a record of unique ways of learning from other humans. Retaining new knowledge, we pass this information to our offspring and other members of our societies, and this pro- cess extends over many generations. That is, social learning makes it possible for humans to accumulate an amazing amount of information over long time periods.

In the following chapters, you will be looking at these processes and behaviors— the particulars of physical anthropology— from a biocultural perspective. It is the unique and phenomenal interplay between biology, culture, and behavior that makes us human.

FIGURE 1.5 Tool- Making Once thought to be a uniquely human phenomenon, tool- making has been observed in chimpanzees, the closest biological relatives of humans. As seen here, chimpanzees have modified twigs to scoop termites from nests. Other chimpanzees have used two rocks as a hammer and anvil to crack open nuts. More recently, gorillas were seen using a stick to test the depth of a pool of water they wanted to cross. Tool use such as this likely preceded the first identified tools (see Figure 1.4).

social learning The capacity to learn from other humans, enabling the accu- mulation of knowledge across many generations.

 

 

14 | CHAPTER 1 What Is Physical Anthropology?

How We Know What We Know: The Scientific Method How do physical anthropologists make decisions about what their subject matter means? More specifically, how do we know what we know about human evolution and human variation? Like all other scientists, physical anthropologists carefully and systematically observe and ask questions about the natural world around them. These observations and questions form the basis for identifying problems and gath- ering evidence— data— that will help answer questions and solve problems— that is, fill gaps in scientific knowledge about how the natural world operates. These data are used to test hypotheses, possible explanations for the processes under study. Scientists observe and then reject or accept these hypotheses. This process of determining whether ideas are right or wrong is called the scientific method (Figure 1.6). It is the foundation of science.

Science (Latin scientia, meaning “knowledge”), then, is more than just knowl- edge of facts about the natural world. Science is also much more than technical details. Certainly, facts and technical details are important in developing answers to questions, but facts and technical skills are not science. Rather, science is a process that provides new discoveries that connect our lives with the world we live in— it is a way of knowing through observation of natural phenomena. This repeated acquisition results in an ever- expanding knowledge base, one built from measurable, repeatable, and highly tangible observations. In this way, science is empirical, or based on observation or experiment. After the systematic collection of observations, the scientist develops a theory— an explanation as to why a natural phenomenon takes place. For many nonscientists, a theory is simply a guess or a hunch; but for a scientist, a theory is not just some stab at an explanation. Rather, a theory is an explanation grounded in a great deal of evidence, or what a lawyer calls the “evidentiary record.” A scientist builds a case by identifying incontrovertible facts. To arrive at these facts, the scientist examines and reexamines the evidence, putting it through many tests.

The scientist thus employs observation, documentation, and testing to gener- ate hypotheses and, eventually, to construct a theory based on those hypotheses. Hypotheses explain observations, predict the results of future investigation, and can be refuted by new evidence.

data Evidence gathered to help answer questions, solve problems, and fill gaps in scientific knowledge.

hypotheses Testable statements that potentially explain specific phenomena observed in the natural world.

scientific method An empirical research method in which data are gathered from observations of phenomena, hypotheses are formulated and tested, and conclu- sions are drawn that validate or modify the original hypotheses.

empirical Verified through observation and experiment.

theory A set of hypotheses that have been rigorously tested and validated, leading to their establishment as a gen- erally accepted explanation of specific phenomena.

Observations

Hypothesis supported Hypothesis rejected

Hypothesis

Predictions (“If… then…”)

Test (observations, experiments)

Further tests

New or revised

hypothesis

FIGURE 1.6 The Scientific Method: How We Know What We Know

 

 

How We Know What We Know: The Scientific Method | 15

For example, the great English naturalist Charles Robert Darwin (1809–1882) (Figure  1.7) developed the hypothesis that the origin of human bipedalism was linked to the shift from life in the trees to life on the ground (Figure 1.8). Darwin’s hypothesis was based on his own observations of humans walking, other scientists’ then- limited observations of nonhuman primate behavior, and other scientists’ anatomical evidence, or information (about structural makeup) drawn from dis- sections, in this case of apes. Darwin’s hypothesis led to an additional hypothesis, based on evidence that accumulated over the following century and a half, that the first hominins arose in the open grasslands of Africa from some apelike animal that was formerly arboreal— that is, had once lived in trees.

Support for Darwin’s hypothesis about human origins— and in particular the origin of bipedal locomotion— began to erode in 2001, when a group of scientists discovered early hominins, from 5.2 to 5.8 mya, in the modern country of Ethiopia. Contrary to expectation and accepted wisdom, these hominins had lived not in grasslands but in woodlands. Moreover, unlike modern humans, whose fingers and toes are straight because we are fully terrestrial— we live on the ground—the early hominins had slightly curved fingers and toes. The physical shape and appearance, what physical anthropologists call morphology, of the hominins’ finger and toe bones indicate a lot of time spent in trees, holding on to branches, moving from limb to limb. These findings forced scientists to reject Darwin’s hypothesis, to toss out what had been a fundamental tenet of physical anthropology.

This story does not end, however, with the understanding that the earliest hominins lived in forests. Instead, this new hypothesis generated new questions. For example, why did the earliest hominins arise in a wooded setting, and why did they “come out of the woods” as time went on? Later in this book, we will consider these questions. For now, the point is that science is a self- correcting approach to knowledge acquisition. Scientists develop new hypotheses as new findings are made. Scientists use these hypotheses to build theories. And like the hypotheses that underlie them, theories can be modified or even replaced by better theories, depending on findings made through meticulous observation. As new observations are made and hypotheses and theories are subjected to the test of time, science revises its own errors.

A scientific law is a statement of irrefutable truth of some action or actions occurring in the natural world. Among the few scientific laws, the well- known ones are the laws of gravity, thermodynamics, and motion. But scientific truth seldom gets finalized into law. Rather, truth is continuously developed— new facts are discovered and new understandings about natural phenomena are made. Unlike theories, scientific laws do not address the larger questions as to why a natural action or actions take place.

FIGURE 1.7 Charles Darwin George Richmond painted this portrait of Darwin in 1840.

anatomical Pertaining to an organism’s physical structure.

arboreal Tree- dwelling, adapted to living in the trees.

terrestrial Life- forms, including humans, that live on land versus living in water or in trees.

morphology Physical shape and appearance.

scientific law A statement of fact describ- ing natural phenomena.

FIGURE 1.8 Bipedalism These 1887 photographs by Eadweard Muybridge capture humans’ habitual upright stance. Other animals, such as chimpanzees, occasionally walk on two feet; but humans alone make bipedalism their main form of locomotion. As Darwin observed, this stance frees the hands to hold objects. What are some other advantages of bipedalism?

 

 

16

As my crew and I traveled to St. Catherines Island, we were intent on discovering new facts and forming new understandings about the prehistoric farmers’ descen- dants who were first encountered by Spaniards in the late 1500s. These facts and understandings would enable us to test hypotheses about human evolution and human variation. Once we completed the months of arduous fieldwork and the years of laboratory investigations on the remains that fieldwork uncovered, we would have some answers. The scientific method would guide us in provid- ing insights into this part of the human lineage— human beings’ most recent evolution— and how our species came to be what it is in the early twenty- first century.

C H A P T E R   1 R E V I E W

A N S W E R I N G T H E B I G Q U E S T I O N S

What is anthropology? • Anthropology is the study of humankind. In two

major ways, it differs from other sciences that study humankind. First, anthropology views humans as both biological and cultural beings. Second, anthropology emphasizes a holistic, comparative approach, encompassing all people at all times and all places.

• The four branches of anthropology are cultural anthropology (study of living cultures), archaeology (study of past cultures), linguistic anthropology (study of language), and physical anthropology.

What is physical anthropology? • Physical (or biological) anthropology is the study

of human biology, specifically of the evolution and variation of humans (and their relatives, past and present).

• Physical anthropology is an eclectic field, deriving theory and method both from within the discipline and from other sciences that address important questions about human evolution and human variation.

What makes us human and different from other animals? • Humans living today are the product of millions of

years of evolutionary history and their own personal life histories.

• Humans have six unique physical and behavioral characteristics: bipedalism, nonhoning chewing, complex material culture and tool use, hunting, speech, and dependence on domesticated foods.

How do physical anthropologists know what they know? • Physical anthropologists derive knowledge via the

scientific method. This method involves observations, the development of questions, and the answering of those questions. Scientists formulate and test hypotheses that they hope will lead to theories about the natural world.

1

2

3

4

REVIEW THIS CHAPTER WITH PERSONALIZED, INTERACTIVE QUESTIONS THROUGH IJK, INQUIZITIVE.WWNORTON.COM q

 

 

17

K E Y T E R M S anatomical anthropology arboreal archaeology artifacts biocultural approach biological anthropology bipedalism cultural anthropology culture

data empirical forensic anthropology genome hominin hypotheses language linguistic anthropology material culture morphology

nonhoning canine physical anthropology primates scientific law scientific method social learning sociolinguistics terrestrial theory

E V O L U T I O N R E V I E W Physical Anthropology as Science

Synopsis Anthropology is a holistic discipline in that it views humankind from the perspectives of all people and all times. Anthropology is also an interdisciplinary science in that it both draws on and influences research in many related fields. Physi- cal anthropology is one of the four subfields (along with cultural anthropology, archaeology, and linguistic anthropology) that make up anthropology as both a biological and a social science. The two main concepts that define physical anthropology are human biological evo- lution and human biocultural variation. Through the employment of the scientific method, physical anthropologists study many different aspects of living humans, modern and extinct nonhuman primates, and fossil hominins, among other lines of research. Through all of these different ways of gathering knowledge about the human con- dition, physical anthropologists ultimately address research questions related to the two broad themes of evolution and variation.

Q1. Define the biocultural approach—a hallmark of physical anthropology.

Q2. Focusing on Figure 1.3, The Six Big Events of Human Evolution, identify which two of these events were caused primarily by biological changes in humans and which four were caused by changes in both human biology and human culture.

Q3. Over time, has culture had more or less of an effect on human evolution? Focusing on Figure 1.3, briefly explain your answer.

Q4 . As a species, humans are unique in the degree to which culture influences our evolution. Consider Figure 1.3 again. How might aspects of human culture have affected the evolution of other species, such as livestock or wild animals?

Q5. Many nonscientists often critique evolution as “just a theory.” What does it mean for evolution to be a theory in the context of the scientific method? How does the study of evolution illustrate the interdisciplinary nature of physical anthropology?

Hint What other scientific fields might contribute data that are used to test hypotheses related to biological evolution?

A D D I T I O N A L R E A D I N G S

Larsen, C. S., ed. 2010. A Companion to Biological Anthropology. Chichester, UK: Wiley- Blackwell.

Molnar,  S.  2005. Human Variation: Races, Types, and Ethnic Groups. Upper Saddle River, NJ: Prentice Hall.

Moore, J. A. 1999. Science as a Way of Knowing: The Foundations of Modern Biology. Cambridge, MA: Harvard University Press.

Spencer,  F., ed. 1997. History of Physical Anthropology: An Ency- clopedia. New York: Garland.

 

 

Stocking,  G., ed. 1974. The Shaping of American Anthropology, 1883–1911: A Franz Boas Reader. New York: Basic Books.

Some Periodicals in Anthropology

Physical anthropology: American Journal of Human Biology, Amer- ican Journal of Physical Anthropology, American Journal of Prima- tology, Evolutionary Anthropology, Human Biology, International

Journal of Paleopathology, Journal of Human Evolution, Yearbook of Physical Anthropology.

Archaeology: American Antiquity, Antiquity, Archaeology, Journal of Archaeological Science, Latin American Antiquity, World Archaeology.

Cultural anthropology: American Anthropologist, Cultural Anthro pology.

General anthropology: American Anthropologist, Annual Review of Anthropology, Current Anthropology.

18

A D D I T I O N A L R E A D I N G S

 

 

19

The Present: Foundation for the Past

Some physical anthropologists learn about human evolution by studying living plants and animals, including humans. Other physical anthropologists learn about human evolution by investigating the past, now represented mostly by fossilized bones and fossilized teeth. Together, living and past enable us to understand evolution in the largest context. The fossil record provides us with the history of humans and of humanlike ancestors, while the living record provides the essential picture through which to view that history. Charles Darwin, the pioneering force behind our knowledge about evolution and natural selection, developed his ideas by studying living plants and animals. He had the extraordinary insight to realize that his

theories and hypotheses applied to past organisms. For Darwin, living organisms were key to interpreting the past because they displayed evidence of evolution’s elements and mechanics. In the same way, living organisms pro- vide insights— into fundamental forces such as reproduc- tion, DNA synthesis, protein synthesis, and behavior— that are not available, at least in the same way, within the past record. Part I of this book lays out observations and princi- ples based on the study of living populations, the essential background for understanding evolution. Part II then digs into the past, into the study of ancestors whose descen- dants are present in the world (all of us now living) and of those evolutionary lineages that did not survive.

The living primates— such as, here, orangutans and humans— have much in common, biological and behavioral. Their study provides essential context for understanding variation and evolution, now and in the past.

P A R T I

 

 

CHARLES DARWIN’S OBSERVATIONS provided the groundwork for his theory of natural selection, the basis of his 1859 book On the Origin of Species.

 

 

21 21

2 How did the theory of evolution come to be?

What was Darwin’s contribution to the theory of evolution?

What has happened since Darwin in the development of our understanding of evolution?

Evolution Constructing a Fundamental Scientific Theory

T he nineteenth century was the century of scientific collecting. During the 1800s, the world discovered itself through collections. Expeditions large and small— involving scientists, explorers, and adventurers— crossed the continents and investigated landmasses around the globe. These teams collected hundreds of thou- sands of samples: plants, animals, rocks, and preserved remains (or fossils— the sub- ject of chapter 8). If it seemed worth picking off the ground or exposing in some other fashion, it was fair game. This kind of work, on one of these international expeditions, helped lay the foundation for the most important biological theory, arguably among the half- dozen most important scientific theories— the theory of evolution.

In 1831, a 22- year- old Englishman and recent graduate of Cambridge University, Charles Darwin, was appointed the naturalist for a five- year voyage around the world on the ship HMS Beagle (Figure 2.1). Imagine that as your first job right out of col- lege! Young Mr. Darwin, who was trained in medicine and theology, accepted a difficult task. He was to collect, document, and study the natural world— plants and animals, especially— everywhere the ship harbored. By the end of that voyage, Darwin had amassed a wonderfully comprehensive collection of plants, insects, birds, shells, fossils, and lots of other materials. The specimens he collected and the observations he made about the things he saw on that trip would form the basis of his lifetime of research. His discoveries would do no less than shape the future of the biological sciences, including physical anthropology. His ideas would provide the key to understanding the origin and evolution of life itself.

Soon after returning home from the voyage, Darwin began to formulate questions about the origins of plants and animals living in the many lands he and his shipmates

1

2

3

B I G Q U E S T I O N S ?

fossils Physical remains of part or all of once- living organisms, mostly bones and teeth, that have become mineralized by the replacement of organic with inorganic materials.

 

 

22 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

had explored. His most prominent observations concerned the physical differences, or variation, between and among members of species, or like animals and like plants. He articulated the phenomenon best in his notes on finches that live in the Galápagos, a small cluster of islands 965 km (600 mi) off the coast of Ecuador. Not only did these birds differ from island to island, but even within a single island they seemed to vary according to habitat, or surroundings. For example, finches living on an island’s coast had a different beak shape from finches living in an island’s inte- rior (Figure 2.2). These observations raised two questions for Darwin: Why were the birds different from island to island and from habitat to habitat? How did differ- ent species of finches arise? After years of study, Darwin answered these questions with an idea called “descent with modification,” or the theory of evolution.

Darwin also came to realize that the variations in physical characteristics of the different species of finches and other organisms were adaptations— physical characteristics that enhance an organism’s ability to survive and reproduce. Dar- win recognized many other adaptations in the natural world, and he concluded that adaptation was the crux of evolution. To connect these processes, he coined the term natural selection. According to this principle, biological characteristics that enhance survival increase in frequency from generation to generation. Members of a population endowed with these characteristics produce more offspring that survive to reproductive age than members that are not endowed with these characteristics. Natural selection is thus the primary driver of evolution. Recognizing that the differ- ent species of finches all derived from a single common ancestor that had originated in South America, Darwin also postulated the process of adaptive radiation: out of one species branch multiple closely related species.

FIGURE 2.1 Darwin’s Voyage (a) Charles Darwin ca. 1855, about 25 years after he set out on HMS Beagle. (b) In this illustration, the ship is passing through the Strait of Magellan, during the South American stretch of (c) its worldwide journey, whose ports of call are here mapped.

(a)

(b)

Galápagos Islands

Western Islands Canary

Islands

Marquesas

Cape Horn

Strait of Magellan

Madagascar Bay of Islands

N O R T H A M E R I C A

B R I T I S H I S L E S

E U R O P E

A F R I C A

A U S T R A L I A

A S I A

S O U T H A M E R I C A

P A C I F I C O C E A N

K I N G G E O R G E ’ S S O U N D

AT L A N T I C O C E A N

Rio de Janeiro

Bahia

Falkland Islands

Montevideo Valparaiso

Hobart

SydneyCape of Good Hope

(c)

species A group of related organisms that can interbreed and produce fertile, viable offspring.

habitat The specific area of the natural environment in which an organism lives.

adaptations Changes in physical struc- ture, function, or behavior that allow an organism or species to survive and repro- duce in a given environment.

natural selection The process by which some organisms, with features that enable them to adapt to the environment, preferentially survive and reproduce, thereby increasing the frequency of those features in the population.

adaptive radiation The diversification of an ancestral group of organisms into new forms that are adapted to specific environmental niches.

 

 

The Theory of Evolution: The Context for Darwin | 23

Darwin regarded evolution as simply biological change from generation to gener- ation. Many evolutionary biologists today limit their definition of evolution to genetic change only. However, nongenetic developmental change— biological change occurring within an individual’s lifetime— can give an adaptive advantage (or disad- vantage) to an individual or individuals within a population. Moreover, genes control developmental processes, which likewise influence other genes.

In subsequent chapters, we will further explore these and other aspects of evolu- tion. Although the core of this book is human evolution or how human biology came to be, understanding human evolution requires understanding the term evolution as it applies to all living organisms. In this chapter, we will take a historical approach to the term and the theory behind it. After reading about its intellectual history before Darwin, Darwin’s contribution, and developments since Darwin, you should have a clear idea of what physical anthropologists and other evolutionary biologists mean by evolution.

The Theory of Evolution: The Context for Darwin Before Darwin’s time, Western scientists’ understanding of Earth and the organ- isms that inhabit it was strongly influenced by religious doctrine. In the Judeo- Christian view, the planet was relatively young, and both its surface and the life- forms on it had not changed since their miraculous creation. By the late 1700s, scientists had realized three key things about the world and its inhabitants: Earth is quite ancient, its surface is very different from what it was in the past, and plants

FIGURE 2.2 Darwin’s Finches Darwin studied the physical variation in finches living on different islands of the Galápagos. Among other attributes, he studied beak shape, which varied from island to island. Eventually, Darwin related each beak shape to diet, especially to the texture of food and how the food was acquired. Finches with larger beaks typically consumed harder foods, such as seeds and nuts, while finches with smaller beaks ate softer foods, such as berries. Darwin concluded that each finch species had adapted to the particular environment and food resources of its island.

 

 

24 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

and animals have changed over time. These realizations about the natural world provided the context for Darwin’s theory of evolution.

To generate his theory, Darwin drew on information from five scientific dis- ciplines: geology, paleontology, taxonomy and systematics, demography, and what is now called evolutionary biology. Geology is the study of Earth, especially with regard to its composition, activity, and history. This discipline has demonstrated the great age of our planet and the development of its landscape. Paleontology is the study of fossils. This discipline has detailed past life- forms, many now extinct. Taxonomy is the classification of past and living life- forms. This disci- pline laid the foundation for systematics, the study of biological relationships over time. Demography is the study of population, especially with regard to birth, survival, and death and the major factors that influence these three key parts of life. Evolutionary biology is the study of organisms and their changes. By investigating the fundamental principles by which evolution operates, Darwin founded this  discipline. In the following sections, we will look at these fields in more detail.

GEOLOGY: RECONSTRUCTING EARTH’S DYNAMIC HISTORY We now know that our planet is 4.6 billion years old and that over time its surface has changed dramatically. If you had espoused these ideas in, say, the late 1600s, you would not have been believed, and you would have been condemned by the Church because you had contradicted the Bible. According to a literal interpre- tation of the Bible, Earth is a few thousand years old and its surface is static. The Scottish scientist James Hutton (1726–1797) became dissatisfied with the biblical interpretation of the planet’s history (Figure 2.3). He devoted his life to studying natural forces, such as wind and rain, and how they affected the landscape in Scotland. Hutton inferred from his observations that these forces changed Earth’s surface in the past just as they do in the present. Wind and rain created erosion, which provided the raw materials— sand, rock, and soil— for the formation of new land surfaces. Over time, these surfaces became stacked one on top of the other, forming layers, or strata, of geologic deposits (Figure  2.4). From the (very long) time it took for these strata to build up, he calculated Earth’s age in the millions of years. This was a revolutionary, indeed heretical, realization.

FIGURE 2.3 James Hutton Hutton (here depicted ca. 1790) founded modern geology with his theory of Earth’s formation. Hutton realized that the same natural processes he observed in Scotland had occurred in the past.

geology The study of Earth’s physical history.

paleontology The study of extinct life- forms through the analysis of fossils.

taxonomy The classification of organisms into a system that reflects degree of relatedness.

systematics The study and classification of living organisms to determine their evo- lutionary relationships with one another.

demography The study of a population’s features and vital statistics, including birth rate, death rate, population size, and population density.

evolutionary biology A specialty within the field of biology; the study of the process of change in organisms.

FIGURE 2.4 Geologic Strata The succession of strata from oldest at the bottom to youngest at the top (as here, in Utah’s Bryce Canyon) marks the formation of new land surfaces over time.

 

 

The Theory of Evolution: The Context for Darwin | 25

Hutton’s idea— that the natural processes operating today are the same as the natural processes that operated in the past— is called uniformitarianism. Few paid much attention to Hutton’s important contribution to our understanding of Earth’s history until the rediscovery of the idea by the Scottish geologist Charles Lyell (1797–1875; Figure  2.5). Lyell devoted considerable energy to thinking and writing about uniformitarianism and its implications for explaining the history of our planet. His calculations of how long it would have taken for all known strata to build up created a mountain of evidence, an undeniable record, that Earth was millions of years old. Hutton and Lyell, relying on empirical evidence and personal observation to develop their ideas and to test clear hypotheses about the natural world, had revised the timescale for the study of past life.

PALEONTOLOGY: RECONSTRUCTING THE HISTORY OF LIFE ON EARTH For hundreds of years, people have been finding the preserved— that is, fossilized— remains of organisms all over the world (see also the full discussion in chapter 8). To test his hypothesis that fossils are the remains of past life, the English scientist Robert Hooke (1635–1703) studied the microscopic structure of fossil wood. After observing that the tissue structure of the fossil wood was identical to the tissue structure of living trees, Hooke concluded that the fossil wood derived from once- living trees (Figure 2.6).

Fossils’ potential to illuminate the past was demonstrated by the French nat- uralist and zoologist Georges Cuvier (1769–1832). Cuvier devoted considerable effort to learning the anatomy, or structural makeup, of many kinds of animals (Figure  2.7). Pioneering what we now call paleontology and comparative anat- omy,  he applied his extensive knowledge of comparative anatomy to fossils. By doing so, he reconstructed the physical characteristics of past animals— their appearance, physiology, and behavior. Although not very accurate by today’s standards, these efforts provided early tools for understanding past life- forms as once- living organisms. Through detailed reconstructions, Cuvier demonstrated

FIGURE 2.5 Charles Lyell Lyell (here depicted ca. 1845) rediscovered Hutton’s work and the idea of uniformitarianism. Lyell’s research, based on examinations of geologic strata, confirmed Hutton’s estimate of Earth’s very old age.

uniformitarianism The theory that pro- cesses that occurred in the geologic past are still at work today.

FIGURE 2.6 Robert Hooke (a) Hooke did pioneering biological research using a very simple microscope. He was the first to identify cells; in fact, he coined the term cell. (b) This illustration of cork wood cells appeared in Hooke’s Micrographia (1667), the first major book on microscopy. His examinations of cells like this enabled Hooke to determine that fossils represented past life- forms.(a) (b)

 

 

26 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

that fossils found in geologic strata in France were the remains of animals that had gone extinct at some point in the remote past. Cuvier’s work provided the first basic understanding of the history of life, from the earliest forms to recent ones.

Cuvier observed that each stratum seemed to contain a unique set of fos- sils.  What happened to the animals represented by each set, each layer? Cuvier concluded that they must have gone extinct due to some powerful catastrophe, such as an earthquake or a volcanic eruption. He surmised that following each catastrophe, the region was vacant of all life and was subsequently repopulated by a different group of animals moving into it from elsewhere. This perspective is called catastrophism.

We now know that Earth’s history does not consist of sequential catastrophes and resulting extinctions. Past catastrophes, such as the extinction of the dinosaurs at around 65 mya, have profoundly affected the direction of evolution, but they were not the leading factor in evolution.

However, such events are rare and do not explain even the sequence of fossils Cuvier observed, mostly in the region called the Paris Basin. In addition to con- firming that fossils were the remains of life in the distant past, though, Cuvier revealed that the most recent geologic strata contain mostly mammals and earlier geologic strata contain mostly reptiles, including the dinosaurs.

TAXONOMY AND SYSTEMATICS: CLASSIFYING LIVING ORGANISMS AND IDENTIFYING THEIR BIOLOGICAL RELATIONSHIPS In the pre- Darwinian world, most scientists who studied life- forms realized the importance of developing a taxonomy— a classification of life- forms— for identifying biological relationships. Early efforts at taxonomy took a commonsense approach. Animals were placed within major groups such as dogs, cats, horses, cattle, and people. Plants were placed within major groups such as trees, shrubs, vines, and weeds.

(b)(a)

FIGURE 2.7 Georges Cuvier (a) One of Cuvier’s most important contributions to science was the concept of extinction. Here, Cuvier is depicted examining a fish fossil. (b) In his 1796 paper on fossil and living elephants, Cuvier suggested that mammoth remains— such as those shown here, from one of his many publications— represented a species different from any living elephant species and, therefore, were from a species that had gone extinct. This idea was revolutionary because the common perception was that God had created all species, none of which had ever gone extinct.

catastrophism The doctrine asserting that cataclysmic events (such as volcanoes, earthquakes, and floods), rather than evolutionary processes, are responsible for geologic changes throughout Earth’s history.

 

 

The Theory of Evolution: The Context for Darwin | 27

As late as the seventeenth century, scientists generally believed that species were immutable. In their view, life had changed very little, or not at all, since the time of the single Creation. Thus, early taxonomists were not motivated by an interest in evolution. Rather, they were motivated by their desire to present the fullest and most accurate picture of the Creator’s intentions for His newly created world. To construct the best possible taxonomy, the English naturalist John Ray (1627–1705) advocated personal observation, careful description, and consideration of plants’ and animals’ many attributes. Ray’s attention to detail laid the

Pre- Darwinian Theory and Ideas: Groundwork for Evolution

Charles Darwin first presented his theory of evolution in his book On the Origin of Species (1859). Based on years of personal observation and of study, this unifying biological theory drew on geology, paleontology, taxonomy and systematics, and demography.

Scientist Contribution (and year of publication) Significance

James Hutton Calculated Earth’s age as millions of years (1788)

Provided geologic evidence necessary for calculating time span of evolution

Charles Lyell Rediscovered and reinforced Hutton’s ideas (1830)

Provided more geologic evidence

Robert Hooke Proved that fossils are organisms’ remains (1665)

Revealed that fossils would provide the history of past life

Georges Cuvier

Extensively studied fossils (1796)

Revealed much variation in the fossil record

John Ray Pioneered taxonomy based on physical appearance (1660)

Created the first scientific classification of plants and animals

Carolus Linnaeus

Wrote Systems of Nature (1735)

Presented the binomial nomenclature taxonomy of plants and animals

Thomas Malthus

Founded demography: only some will find enough food to survive (1798)

Provided the concept of characteristics advantageous for survival

Jean- Baptiste de Lamarck

Posited characteristics acquired via inheritance (Lamarckism) (1809)

Provided first serious model of physical traits’ passing from parents to offspring

Erasmus Darwin

Also posited characteristics (determined by wants and needs) acquired via inheritance (1794)

Advanced the notion that physical changes occurred in the past

C O N C E P T C H E C K !

 

 

28 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

groundwork for later taxonomy, especially for the binomial nomenclature ( two- name) system developed by the Swedish naturalist Carl von Linné (1707–1778). Von Linné, better known by his Latinized name, Carolus Linnaeus, gave each plant and animal a higher- level genus (plural, genera) name and a lower- level species (plural is also species) name (Figure 2.8). A single genus could include one or more spe- cies. For example, when Linnaeus named human beings Homo sapiens— Homo being the genus, sapiens being the species— he thought there were species and subspecies of living humans (an idea discussed further in chapter  5). The presence of more than one level in his taxonomy acknowledged different degrees of physical simi- larity. Today, we recognize that sapiens is the one living species in the genus Homo.

Linnaeus presented the first version of his taxonomy in his book Systema Naturae (1735), or System of Nature. As he revised the taxonomy— his book would eventually go through 10 editions— he added more and more levels to the hierarchy. He clas- sified groups of genera into orders and groups of orders into classes. For example, he named the order “Primates,” the group of mammals that includes humans, apes, monkeys, and prosimians. Since the eighteenth century, this taxonomic system has evolved into multiple levels of classification, going from the subspecies at the bottom to the kingdom at the top (Figure 2.9).

Like Ray, Linnaeus was committed to the notion that life- forms were static, fixed at the time of the Creation. In later editions of his book, he hinted at the pos- sibility that some species may be related to each other because of common descent, but he never developed these ideas. His taxonomy is still used today, although viewed with a much stronger sense of present and past variation. The system’s flexibility aided evolutionary biologists in their study of biological diversity, and the focus on taxonomic relationships over time is now called systematics.

DEMOGRAPHY: INFLUENCES ON POPULATION SIZE AND COMPETITION FOR LIMITED RESOURCES After returning to England and while developing his ideas on natural selection, Darwin read the works of all the great scientists of the time. Probably the most important influence on his ideas was An Essay on the Principle of Population, by the English political economist Thomas Malthus (1766–1834). First published in 1798, Malthus’s book made the case that an abundance of food— enough to feed anyone born— would allow the human population to increase geometrically and indefinitely. In reality, the Essay argued, there simply is not enough food for every- one born, so population is limited by food supply (Figure 2.10). Who survives to reproductive age? Those who can successfully compete for food. Whose children thrive? Those of survivors who manage to feed their offspring. Applying Malthus’s demographic ideas to human and nonhuman animals, Darwin concluded that some members of any species successfully compete for food because they have some special attribute or attributes. That an individual characteristic could facilitate survival was a revelation!

EVOLUTIONARY BIOLOGY: EXPLAINING THE TRANSFORMATION OF EARLIER LIFE- FORMS INTO LATER LIFE- FORMS By the late 1700s, a handful of scientists had begun to argue that, contrary to religious doctrine, organisms are not fixed— they change over time, sometimes in dramatic ways. Simply, life evolved in the past and evolution is an ongoing,

genus A group of related species.

FIGURE 2.8 Carolus Linnaeus Linnaeus, a botanist, zoologist, and physician, is known for his contributions to the system of classification used today by all biological scientists, including physical anthropologists. He is also a founder of modern ecology.

FIGURE 2.10 Thomas Malthus Malthus, the founder of demography, theorized that population size was limited by food supply.

 

 

TAXONOMIC CATEGORY

Subfamily Homininae

Homo

sapiens

sapiens Modern humans alone.

Tribe Hominini

Genus

Species

Subspecies

TAXONOMIC LEVEL COMMON CHARACTERISTICS

Group of hominins including modern humans, their direct ancestors, and extinct relatives (e.g., Neandertals). They have the largest brains in the Hominini.

Modern and ancestral modern humans. They have culture, use language, and inhabit every continent except Antarctica.

Chimpanzees, humans, and humanlike ancestors

Family Hominidae Great apes, humans, and humanlike ancestors.

Superfamily Hominoidea Group of anthropoids, including humans, great apes, lesser apes, and humanlike ancestors. They have the largest bodies and brains of all primates.

Parvorder Catarrhini Group of anthropoids, including humans, apes, and Old World monkeys.

Infraorder Anthropoidea Group of haplorhines, including humans, apes, and monkeys.

Suborder

Order Primates

Haplorhini Group of primates, including monkeys, apes, humans, and tarsiers. They have in general long life cycles and are relatively large-bodied.

Group of mammals specialized for life in the trees, with large brains, stereoscopic vision, opposable thumbs, and grasping hands and feet.

Subclass Theria Group of mammals that produce live young without a shelled egg (including placental and marsupial mammals).

Class Mammalia Group of warm-blooded vertebrate animals that produce milk for their young in mammary glands. They have hair or fur and specialized teeth.

Superclass Tetrapoda Vertebrate animals with four feet or legs, including amphibians, birds, dinosaurs, and mammals.

Subphylum Vertebrata Animals with vertebral columns or backbones (including fish, amphibians, reptiles, birds, and mammals).

Phylum Chordata Group of vertebrate and invertebrate animals that have a notochord, which becomes the vertebral column in humans and other primates.

Subkingdom Eumetazoa All major animals (except sponges) that contain true tissue layers, organized as germ layers, which develop into organs in humans.

Kingdom Animalia Mobile multicellular organisms that consume other organims for food and develop during an embryo stage.

Commonly called “hominins,” this level includes humans and humanlike ancestors, all of which are obligate bipeds.

FIGURE 2.9 The Place of Humans in Linnaeus’s Taxonomy Linnaeus’s system organized living things into various levels of hierarchical classification. Kingdom, at the top of the taxonomy, is the largest classification. The five kingdoms of the natural world— animals, plants, fungi, protists, monera— include all living organisms. Through descending taxonomic levels, each group’s size gets progressively smaller. For example, there are fewer organisms in a genus than there are in a phylum. Additionally, these classifications reflect organisms’ relationships to one another. For example, organisms within a genus are more closely related than are those from different genera.

 

 

30 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

undirected process. Building on this concept, the French naturalist Jean- Baptiste de Monet (1744–1829), better known by his title, Chevalier de Lamarck, speculated that plants and animals not only change in form over time but do so for purposes of self- improvement. Lamarck believed that in response to new demands or needs, life- forms develop new anatomical modifications, such as new organs. His central idea— that when life- forms reproduce, they pass on to their offspring the modifications they have acquired to that point— is called Lamarckian inheritance of acquired characteristics, or Lamarckism (Figure  2.11). We now know Lamarck’s mechanism for evolution to be wrong— offspring do not inherit traits acquired by their parents— but his work was the first major attempt to develop a theory built on the premise that living organisms arose from precursor species. Lamarck was also convinced that humans evolved from some apelike animal.

Among the other scholars who believed that life had changed over time was the English physician, naturalist, and poet Erasmus Darwin (1731–1802), grandfather of Charles Darwin. Like Lamarck, he hypothesized about the inheritance of char- acteristics acquired thanks to wants and needs; but he, too, was wrong about the mechanism for change.

Lamarckism First proposed by Lamarck, the theory of evolution through the inheritance of acquired characteristics in which an organism can pass on features acquired during its lifetime.

Darwin Borrows from Malthus

Five of Malthus’s observations inspired Darwin’s principle of natural selection.

Observation 1 For most organisms, every pair of parents produces multiple (sometimes many) offspring.

Observation 2 For most organisms, the population size remains the same. No increase occurs over time.

Observation 3 Population is limited by the food supply.

Observation 4 Members of populations compete for access to food.

Observation 5 No two members of a species are alike in their physical attributes— variation exists.

Theory: Evolution by Means of Natural Selection Individuals having variation that is advantageous for survival to reproductive age produce more offspring (and more offspring that survive) than individuals lacking this variation.

C O N C E P T C H E C K !

 

 

The Theory of Evolution: Darwin’s Contribution | 31

The Theory of Evolution: Darwin’s Contribution Darwin’s remarkable attention to detail enabled him to connect his voluminous reading with his personal observations from the Beagle voyage. For example, while in Chile, Darwin had observed firsthand the power of earthquakes in shaping the

(a)

Original short-necked ancestor.

Descendants keep stretching to reach leaves higher up on tree …

… and stretching until neck becomes progressively longer in descendants.

… and stretching …

(b)

FIGURE 2.11 Jean- Baptiste Lamarck (a) Lamarck developed an early theory of evolution involving the inheritance of acquired characteristics. Although his mechanism of evolution was wrong, Lamarck’s recognition of the dynamic nature of life in the past made an important contribution to the development of evolutionary theory. (b) According to the classic (though incorrect) example of Lamarckism, giraffes stretched to reach food at the tops of trees, their necks grew as a result, and they passed on these long necks to their offspring.

 

 

32 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

landscape. Hutton’s and Lyell’s uniformitarianism led him to recognize that the accumulation of such catastrophes over a long period of time explains, at least in part, the appearance of the present- day landscape. This understanding of Earth’s remarkably dynamic geologic history laid the groundwork for Darwin’s view of evolution as a long, gradual process.

That process, he saw, could be reconstructed through the fossil record. He had read carefully Cuvier’s studies of fossils, and in South America he saw fossils first- hand. Some of these fossils resembled living animals native to South America, such as the armadillo, ground sloth, and llama. This evidence strongly suggested that an earlier species had transformed into the modern species, most likely through a succession of species over time. Drawing on Malthus’s ideas about reproduction, population, and variation, Darwin wrote, “it at once struck me that under these circumstances [i.e., specific environmental conditions] favorable variations would

(a)

(b)

FIGURE 2.12 Writing a Masterpiece (a) Darwin wrote most of On the Origin of Species at his beloved home, Down House, in Kent, England. (b) He generally worked in his study there.

 

 

Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA | 33

tend to be preserved and unfavorable ones to be destroyed. The result of this would be the formation of new species.” Another revelation.

Darwin hypothesized that surviving offspring had attributes advantageous for acquiring food. Because these offspring survived, the frequency of their advanta- geous characteristics increased over time. Meanwhile, as environmental conditions changed— such as when rainfall decreased— offspring lacking adaptive attributes suited to their survival in the new environment died off. Building on these obser- vations and their implications, Darwin deduced that natural selection was the primary mechanism of evolution. Over a long period of time, through generations’ adaptation to different environments and different foods, a common ancestor gave rise to related species. Darwin’s hypothesis was revolutionary, undermining the mid- nineteenth- century consensus that species were fixed types in a defined natural order of life. Now, species would have to be considered as populations with no predetermined limit on variation.

Darwin’s background research had begun in the 1830s. It was not until 1856, however— fully two decades after his voyage around the world— that Darwin had gathered enough evidence and developed his ideas enough to begin writing his great work about evolution by means of natural selection, On the Origin of Species (Figure 2.12). His colleagues had warned him that if he did not write his book soon, someone else might receive credit for the idea. Indeed, in 1858, Darwin received from the English naturalist and explorer Alfred Russel Wallace (1823–1913) a letter and a 20-page report outlining Wallace’s theory of evolution by means of natural selection (Figure 2.13). Independently from Darwin, Wallace had arrived at most of the same conclusions that Darwin had. Both men had been aware of their shared interest in the subject, and both men formulated their theories independently. Concerned that Wallace would publish first, Darwin completed Origin over the next 15 months and published it in London in 1859. Who, then, “discovered” natu- ral selection, the key mechanism that explains evolution? Some argue that Wallace should be given primary credit for the theory. However, because Wallace had not amassed the extensive body of evidence needed to support the theory, Darwin is generally recognized as the discoverer.

Darwin and Wallace made monumental discoveries, but neither man could come up with a compelling explanation of the physical mechanisms by which evolution takes place. That is, what mechanisms result in evolutionary change? Half a continent away, a series of novel experiments led to the discovery of these biological mechanisms, paving the way for remarkable new insights into evolution.

Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA MECHANISMS OF INHERITANCE Having articulated and supported his theory of evolution by means of natural selection, Darwin turned to the next fundamental question about natural selec- tion: How do the traits that are being selected for (or against) pass from parent to offspring? Like other scientists of his day, Darwin believed that each body part contained invisible particles called gemmules. Darwin hypothesized that

FIGURE 2.13 Alfred Russel Wallace Although Darwin often gets sole credit for the development of the theory of evolution through natural selection, Wallace (here depicted ca. 1860) contributed substantially to evolutionary theory. Wallace was the leading authority on the geographic distribution of animals, for example, and was the first to recognize the concept of warning coloration in animals. In addition, he raised the issue of human impact on the environment a full century before it became a concern for the general public.

gemmules As proposed by Darwin, the units of inheritance, supposedly accu- mulated in the gametes so they could be passed on to offspring.

 

 

34 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

representative gemmules for all body parts resided in the reproductive organs. During fertilization, each parent contributed his or her gemmules to the poten- tial offspring. The father’s and the mother’s gemmules then intermingled to form the characteristics observed in their progeny. Called blending inheritance, this process was a popular notion at the time.

Unknown to Darwin, research elsewhere in Europe was calling into question the idea of blending inheritance. In 1865, just six years after the publication of On the Origin of Species, Gregor Mendel (1822–1884), an Augustinian monk living in a mon- astery in what is now Brno, Czech Republic, published in an obscure local scientific journal the results of his work on inheritance (Figure 2.14). Mendel had spent the previous eight years crossbreeding different varieties of garden pea plants. Over the course of his experiments, he grew some 28,000 plants. These plants enabled him to identify and carefully observe seven characteristics, or traits, that were especially informative about breeding and its outcome over generations (Figure 2.15). From his results, Mendel inferred that a discrete physical unit was responsible for each characteristic. This unit passed from parent to offspring, and in this way the charac- teristic was inherited. In fact, the discrete unit could be traced through generations, and its passage (the inheritance) was determined by mathematical laws.

Mendel also discovered that the garden peas’ traits did not blend. For example, plants and their offspring were either tall or short. Over time, the short plants diminished in frequency and eventually disappeared. Later scientists determined that the physical unit of inheritance— now known as a gene— has two subunits, one from the father and one from the mother, each called an allele. Each allele is either dominant or recessive. In garden peas, the allele for tallness is dominant and the allele for shortness is recessive. If one parent provides a “tall” allele (T ) and the other parent provides a “short” allele (t), then the offspring having one of each allele (Tt) would be tall because of the presence of the “tall” allele— the dominant allele is physically expressed, whereas the recessive allele is hidden. The pure strain for tall (TT ) includes one tall maternal allele (T ) and one tall paternal allele (T ). The pure strain for short (tt) includes one short maternal allele (t) and one short paternal allele (t) (Figure 2.16).

FIGURE 2.14 Gregor Mendel Mendel, the father of modern genetics, was a Christian monk by profession but a scientist by nature. His observations provided the foundation for our understanding of genetics (the subject of chapters 3 and 4).

FIGURE 2.15 Mendel’s Peas (a) This illustration— from the 1876 catalog of one of Mendel’s seed suppliers— shows some of (b) the seven characteristics Mendel studied, each of which had two variants. Flower position, for example, could be axial or terminal, while flower color could be white or purple.

Flower Position Flower Color Plant Height Pea Shape Pea Color Pod Shape Pod Color

Axial White Tall Round Yellow In ated Yellow

Terminal Purple Short Wrinkled Green Constricted Green

(a) (b)

Parent 1 (TT ): Tall

Generation 1

Parent 2 (tt): Short

100% Tt = Tall

t

T

25% TT = Tall 50% Tt = Tall 25% tt = Short

3:1 Tall:Short

t

Tt Tt

T

Tt Tt

Parent 1 (Tt): Tall

Generation 2

Parent 2 (Tt): Tall

T

T

t Tt tt

t

TT Tt

If the tallness allele is expressed as T and the shortness allele is expressed as t, the pure strain for tall is TT (one T is the maternal allele, the other T is the paternal allele), and the pure strain for short is tt (one t is the maternal allele, and the other t is the paternal allele).

When a TT plant is crossbred with a tt plant, one allele must come from the father (paternal) and one allele (maternal) must come from the mother, thereby producing a Tt offspring.

When the offspring from the TT and tt parental plants are bred, the offspring’s alleles independently redistribute, producing about equal numbers of the four possible combinations of T and t alleles: TT, Tt, tT, and tt.

Thus, three of the four plants (75%) will be tall owing to the dominance of the T allele and one plant (25%) will be short owing to the recessiveness of the t allele. Note, however, that 25% of the offspring are tall with two dominant tall alleles (TT ), while 50% are tall with one of each allele (Tt).

Because T is dominant, the offspring is tall.

FIGURE 2.16 Mendel’s Genetics

blending inheritance An outdated, dis- reputed theory that the phenotype of an offspring was a uniform blend of the parents’ phenotypes.

gene The basic unit of inheritance; a sequence of DNA on a chromosome, coded to produce a specific protein.

allele One or more alternative forms of a gene.

dominant Refers to an allele that is expressed in an organism’s phenotype and that simultaneously masks the effects of another allele, if another one is present.

recessive An allele that is expressed in an organism’s phenotype if two copies are present but is masked if the dominant allele is present.

 

 

Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA | 35

representative gemmules for all body parts resided in the reproductive organs. During fertilization, each parent contributed his or her gemmules to the poten- tial offspring. The father’s and the mother’s gemmules then intermingled to form the characteristics observed in their progeny. Called blending inheritance, this process was a popular notion at the time.

Unknown to Darwin, research elsewhere in Europe was calling into question the idea of blending inheritance. In 1865, just six years after the publication of On the Origin of Species, Gregor Mendel (1822–1884), an Augustinian monk living in a mon- astery in what is now Brno, Czech Republic, published in an obscure local scientific journal the results of his work on inheritance (Figure 2.14). Mendel had spent the previous eight years crossbreeding different varieties of garden pea plants. Over the course of his experiments, he grew some 28,000 plants. These plants enabled him to identify and carefully observe seven characteristics, or traits, that were especially informative about breeding and its outcome over generations (Figure 2.15). From his results, Mendel inferred that a discrete physical unit was responsible for each characteristic. This unit passed from parent to offspring, and in this way the charac- teristic was inherited. In fact, the discrete unit could be traced through generations, and its passage (the inheritance) was determined by mathematical laws.

Mendel also discovered that the garden peas’ traits did not blend. For example, plants and their offspring were either tall or short. Over time, the short plants diminished in frequency and eventually disappeared. Later scientists determined that the physical unit of inheritance— now known as a gene— has two subunits, one from the father and one from the mother, each called an allele. Each allele is either dominant or recessive. In garden peas, the allele for tallness is dominant and the allele for shortness is recessive. If one parent provides a “tall” allele (T ) and the other parent provides a “short” allele (t), then the offspring having one of each allele (Tt) would be tall because of the presence of the “tall” allele— the dominant allele is physically expressed, whereas the recessive allele is hidden. The pure strain for tall (TT ) includes one tall maternal allele (T ) and one tall paternal allele (T ). The pure strain for short (tt) includes one short maternal allele (t) and one short paternal allele (t) (Figure 2.16).

FIGURE 2.14 Gregor Mendel Mendel, the father of modern genetics, was a Christian monk by profession but a scientist by nature. His observations provided the foundation for our understanding of genetics (the subject of chapters 3 and 4).

FIGURE 2.15 Mendel’s Peas (a) This illustration— from the 1876 catalog of one of Mendel’s seed suppliers— shows some of (b) the seven characteristics Mendel studied, each of which had two variants. Flower position, for example, could be axial or terminal, while flower color could be white or purple.

Flower Position Flower Color Plant Height Pea Shape Pea Color Pod Shape Pod Color

Axial White Tall Round Yellow In ated Yellow

Terminal Purple Short Wrinkled Green Constricted Green

(a) (b)

Parent 1 (TT ): Tall

Generation 1

Parent 2 (tt): Short

100% Tt = Tall

t

T

25% TT = Tall 50% Tt = Tall 25% tt = Short

3:1 Tall:Short

t

Tt Tt

T

Tt Tt

Parent 1 (Tt): Tall

Generation 2

Parent 2 (Tt): Tall

T

T

t Tt tt

t

TT Tt

If the tallness allele is expressed as T and the shortness allele is expressed as t, the pure strain for tall is TT (one T is the maternal allele, the other T is the paternal allele), and the pure strain for short is tt (one t is the maternal allele, and the other t is the paternal allele).

When a TT plant is crossbred with a tt plant, one allele must come from the father (paternal) and one allele (maternal) must come from the mother, thereby producing a Tt offspring.

When the offspring from the TT and tt parental plants are bred, the offspring’s alleles independently redistribute, producing about equal numbers of the four possible combinations of T and t alleles: TT, Tt, tT, and tt.

Thus, three of the four plants (75%) will be tall owing to the dominance of the T allele and one plant (25%) will be short owing to the recessiveness of the t allele. Note, however, that 25% of the offspring are tall with two dominant tall alleles (TT ), while 50% are tall with one of each allele (Tt).

Because T is dominant, the offspring is tall.

FIGURE 2.16 Mendel’s Genetics

blending inheritance An outdated, dis- reputed theory that the phenotype of an offspring was a uniform blend of the parents’ phenotypes.

gene The basic unit of inheritance; a sequence of DNA on a chromosome, coded to produce a specific protein.

allele One or more alternative forms of a gene.

dominant Refers to an allele that is expressed in an organism’s phenotype and that simultaneously masks the effects of another allele, if another one is present.

recessive An allele that is expressed in an organism’s phenotype if two copies are present but is masked if the dominant allele is present.

 

 

36 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

While Darwin’s theory generated immediate excitement in the scientific com- munity and among the public and was supported by leading scientists of the time such as Thomas Henry Huxley (Figure  2.17), Mendel’s crucial discovery (now known as Mendelian inheritance) went unnoticed. His writing was not widely distributed, and his work was simply ahead of its time. But in 1900, three scientists working independently— the German botanist Carl Erich Correns (1864–1933), the Austrian botanist Erich Tschermak von Seysenegg (1871–1962), and the Dutch botanist Hugo de Vries (1848–1935)—discovered Mendel’s research and replicated his findings. The Danish botanist Wilhelm Ludvig Johannsen (1857–1927) called the pair of alleles (e.g., TT, Tt, tt) the genotype and the actual physical appearance (tall, short) the phenotype.

Mendel’s theory of inheritance forms the basis of the modern discipline of genetics (the subject of chapters 3 and 4). It makes clear that the physical units— the genes and the two component alleles of each gene— responsible for physical attributes are located in the reproductive cells, eggs and sperm. When microscope technology improved in the late nineteenth century, the cell structure and the units of inheritance were defined (see chapter 3).

Beginning in 1908, the American geneticist Thomas Hunt Morgan (1866–1945) and his associates bred the common fruit fly in experiments that built on Mendel’s pea breeding. All genes, they discovered, are transmitted from parents to offspring in the ratios identified by Mendel. The genes are on chromosomes, and both the hereditary material and its carriers are duplicated during reproductive cell division.

THE EVOLUTIONARY SYNTHESIS, THE STUDY OF POPULATIONS, AND THE CAUSES OF EVOLUTION The combination of Darwin’s theory of evolution and Mendel’s theory of heredity resulted in an evolutionary synthesis. Darwin’s theory provided the mechanism for evolution (natural selection), and Mendel’s theory showed how traits are passed on systematically and predictably (Mendelian inheritance). The melding of natural selection and Mendelian inheritance led biologists to ask further questions about evolution, specifically about the origins of particular genes, genetic variation in general, and change in physical characteristics over time. Why do some genes increase in frequency, some decrease in frequency, and some show no change? How do completely new genes appear? These questions and a focus on population— viewed as the gene pool— provided the basis for a newly emerging field in evolution- ary biology called population genetics (among the subjects of chapter 4).

Natural selection, the guiding force of evolution, could operate only on variation that already existed in a population. How did new variation— new characteristics— arise in a population? Through his experiments with fruit flies, Morgan showed that a new gene could appear as a result of spontaneous change in an existing gene. This kind of genetic change is called mutation (Figure 2.18). The only source of new genetic material, mutation is another cause of evolution.

Gene flow, a third cause of evolution, is the diffusion, or spread, of new genetic material from one population to another of the same species. In other words, via reproduction, genes from one gene pool are transferred to another gene pool. Take, for example, the gene that causes sickle- cell anemia (this disorder is discussed extensively in chapter 4). Among West African blacks, it has a frequency of about 10%. Among American whites, it has a frequency of 0%. Because West African blacks and their descendants have long reproduced with American whites, the

FIGURE 2.17 Thomas Henry Huxley Huxley (1825–1895), an English biologist, was known as “Darwin’s bulldog” because he so forcefully promoted Darwin’s theory of evolution by natural selection. Among Huxley’s contributions to evolutionary theory was the concept that humans evolved from an apelike animal.

Mendelian inheritance The basic principles associated with the transmission of genetic material, forming the basis of genetics, including the law of segregation and the law of independent assortment.

genotype The genetic makeup of an organism; the combination of alleles for a given gene.

phenotype The physical expression of the genotype; it may be influenced by the environment.

chromosomes The strand of DNA found in the nucleus of eukaryotes that contains hundreds or thousands of genes.

evolutionary synthesis A unified theory of evolution that combines genetics with natural selection.

population genetics A specialty within the field of genetics; it focuses on the changes in gene frequencies and the effects of those changes on adaptation and evolution.

 

 

frequency among people descended from both West African blacks and American whites of the gene that causes sickle- cell anemia is approximately 5%, halfway between that of the two original populations. Over time, as the two populations have mixed, gene flow has decreased genetic difference.

Genetic drift, the fourth cause of evolution, is random change in the frequency of alleles— that is, of the different forms of a gene. Such change affects a small population more powerfully than it affects a large population (Figure 2.19). Over time, it increases the genetic difference between two genetically related but not interbreeding populations.

By the mid- twentieth century, the four causes of evolution— natural selection, mutation, gene flow, and genetic drift— were well defined, thanks to a synthesis of ideas drawn from the full range of sciences that deal with biological variation. In effect, evolutionary synthesis unified the branches of biology and its affiliated sci- ences, including genetics, taxonomy, morphology, comparative anatomy, paleontol- ogy, and the subject of this book, physical anthropology. Similarly, evolution unites living and past worlds. All organisms are related through common descent, and organisms more closely related than others share a more recent common ancestor.

DNA: DISCOVERY OF THE MOLECULAR BASIS OF EVOLUTION Once chromosomes were recognized as the carriers of genes, scientists sought to understand the structure of deoxyribonucleic acid (DNA), the chemical that makes up chromosomes. In 1953, the American geneticist James Watson (b. 1928) and the British biophysicist Francis Crick (1916–2004) published their discovery that DNA molecules have a ladderlike, double- helix structure. Crucial to their discovery was the work of the British X- ray crystallographer Rosalind Franklin (1920–1958), who used a special technique, X- ray diffraction, to produce high- quality images of  DNA.  The combined efforts of Franklin, Watson, and Crick opened up a whole new vista for biology by helping explain how chromosomes are replicated.

Analysis of the DNA from a wide variety of organisms, including primates, has provided both new perspectives on biological relationships and a molecular “clock” with which to time the branches of evolution (based on the similarity of species within those branches). In addition, DNA analysis has begun to shed light on a growing list of illnesses such as viral and bacterial infections, cancer, heart disease, and stroke.

mutation A random change in a gene or chromosome, creating a new trait that may be advantageous, deleterious, or neutral in its effects on the organism.

gene flow Admixture, or the exchange of alleles between two populations.

genetic drift The random change in allele frequency from one generation to the next, with greater effect in small populations.

deoxyribonucleic acid (DNA) A double- stranded molecule that provides the genetic code for an organism, consisting of phosphate, deoxyribose sugar, and four types of nitrogen bases.

(a) (b)

FIGURE 2.18 Fruit Fly Mutations (a) The normal fruit fly has two wings, while (b) the four- wing mutation has two wings on each side.

Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA | 37

 

 

38 | CHAPTER 2 Evolution: Constructing a Fundamental Scientific Theory

Little did Darwin realize just what a powerful foundation his evolutionary theory would build for science, ushering in modern biology and its allied disciplines, including physical anthropology. Long after his death, Darwin’s search for the biological mechanisms involved in evolution would continue to inspire scientists. The questions Darwin and his colleagues asked, especially about how physical attributes pass from parents to offspring, laid the foundation for the study of inheritance— the science of genetics— and eventually the DNA revolution. Darwin would have been impressed.

One month later

One red fish lost from each population

1:5 8:16

A B A B

0:5 7:16

16.6% Red 33.3% Red 0% Red 32.8% Red

Two populations of fish include red and gold varieties.

In the smaller population (6 fish), the ratio of red to gold is 1:5. Red fish represent 16.6% of the total.

In the smaller population (now 5 fish), the ratio of red to gold changes to 0:5. Red fish represent 0% of the total, a substantial change in the makeup of the population.

In the larger population (24 fish), the ratio of red fish to gold fish is 8:16 or 1:2. Red fish represent 33.3% of the total.

Over time, each population loses one red fish.

In the larger population (now 23 fish), the ratio of red to gold changes to 7:16. Red fish represent 32.8% of the total, a very small change in the makeup of the population.

FIGURE 2.19 Genetic Drift’s Effects on Small and Large Populations

 

 

39

A N S W E R I N G T H E B I G Q U E S T I O N S

C H A P T E R   2 R E V I E W

How did the theory of evolution come to be? • In developing his theory of evolution by means

of natural selection, Darwin drew on geology, paleontology, taxonomy and systematics, demography, and what is now called evolutionary biology.

• Scientists working in these disciplines had shown that — Earth is quite old and has changed considerably

over its history — fossils represent the remains of once- living, often

extinct organisms and thus provide a record of the history of life on the planet

— life evolves over time — groups of related species provide insight into

evolutionary history — the number of adults in a population tends to

remain the same over time

What was Darwin’s contribution to the theory of evolution? • Darwin’s key contribution was the principle of natural

selection. Three observations and inferences allowed him to deduce that natural selection is the primary driver of evolution: — the number of adults in a population tends to

remain the same over time even though, for most

organisms, parents tend to produce multiple and sometimes many offspring

— variation exists among members of populations — individuals having variation that is advantageous

for survival and reproduction increase in relative frequency over time

What has happened since Darwin in the development of our understanding of evolution? • Gregor Mendel discovered the principles of

inheritance, the basis for our understanding of how physical attributes are passed from parents to offspring.

• Mendel’s revelation that attributes are passed as discrete units, which we now know as genes, laid the groundwork for our understanding of cell biology, our understanding of chromosomes, and eventually the field of population genetics.

• We now know that evolution— genetic change in a population or species— is caused by one or a combination of four forces: natural selection, mutation, gene flow, and genetic drift.

• We now know that each chromosome in an organism’s cells consists of DNA molecules. DNA is the blueprint for all biological characteristics and functions.

1

2

3

REVIEW THIS CHAPTER WITH PERSONALIZED, INTERACTIVE QUESTIONS THROUGH IJK, INQUIZITIVE.WWNORTON.COM q

K E Y T E R M S adaptations adaptive radiation allele

blending inheritance catastrophism chromosomes

demography deoxyribonucleic acid (DNA) dominant

 

 

40

K E Y T E R M S

evolutionary biology evolutionary synthesis fossils gemmules gene gene flow genetic drift genotype

genus geology habitat Lamarckism Mendelian inheritance mutation natural selection paleontology

phenotype population genetics recessive species systematics taxonomy uniformitarianism

E V O L U T I O N R E V I E W Past, Present, and Future of a Fundamental Scientific Theory

Synopsis The theory of evolution forms the foundation of all the biological sciences, including physical anthropology. Although Charles Darwin is the most famous contributor to the formulation of this theory, his innovative idea of natural selection was partly influenced by the work of scientists across a number of disciplines, including geology, paleontology, taxonomy, and demography. The work of Gregor Mendel, rediscovered years after his death, provided a genetic basis for the evolutionary processes envisioned by Darwin and showed how evolution can occur in the natural world. Darwin’s principle of natural selection and Mendel’s principles of inheritance are intertwined in the modern evolutionary synthesis, the frame- work by which physical anthropologists address research questions related to human biological evolution and biocultural variation.

Q1. Charles Darwin is one of the most admired scientists of all time, and his principle of natural selection laid the foundations for all future biological thinking and discoveries. However, other scientists before Darwin argued in favor of biological evolution. Who is credited with one of the first major attempts at explaining the process of evolutionary change through time? What is the name of the erroneous mechanism hypothesized by this scientist to be a driving force of evolution?

Q2. Before the discoveries of Gregor Mendel, Darwin hypothesized that the characteristics of the father and mother intermingled

in the offspring. What was this idea called at the time? What  discovery, first made by Mendel and later by scientists such as Thomas Hunt Morgan, proved this hypothesis to be wrong?

Q3. In formulating his principle of natural selection, Darwin was inspired by the idea of the demographer Thomas Malthus that population is limited by food supply. How is this idea a concern for human populations today? What steps might be taken to address this issue in the future?

Q4 . Darwin originally did not publish his theory of evolution by means of natural selection as he was well aware of the con- troversy it would generate. Over 150 years later, and backed by massive amounts of evidence spanning many scientific disciplines, evolution remains a subject of controversy among the general public. Why has evolution always been the subject of fierce debate?

Q5. Darwin gathered information from the fields of geology, paleontology, taxonomy, demography, and what is now called evolutionary biology to develop his theory of evolution, which includes the idea of variation and natural selection. What are the five most important ideas from these other fields (described in this chapter) that contributed to Darwin’s devel- opment of his theory of evolution?

 

 

41

A D D I T I O N A L R E A D I N G S

Alvarez, W. 1997. T. rex and the Crater of Doom. Princeton: Prince- ton University Press.

Berra, T. M. 2009. Charles Darwin: The Concise Story of an Extraor- dinary Man. Baltimore: Johns Hopkins University Press.

Bowler,  P.  J.  2003. Evolution: The History of an Idea. Berkeley: University of California Press.

Carroll,  S.  B.  2009. Remarkable Creatures: Epic Adventures in the Search for the Origins of Species. Boston: Houghton Mifflin Harcourt.

Gould, S. J. 1992. Ever since Darwin: Reflections on Natural History. New York: Norton.

Huxley,  R.  2007. The Great Naturalists. New  York: Thames & Hudson.

Repcheck, J. 2003. The Man Who Found Time: James Hutton and the Discovery of the Earth’s Antiquity. Cambridge, MA: Perseus Publishing.

Ridley, M. 2004. Evolution. Malden, MA: Blackwell Science.

Stott,  R.  2012. Darwin’s Ghosts: The Secret History of Evolution. New York: Spiegel & Grau.

Wilson,  E.  O.  2006. From So Simple a Beginning: Darwin’s Four Great Books [Voyage of the H.M.S. Beagle, The Origin of Species, The Descent of Man, The Expression of Emotions in Man and Ani- mals]. New York: Norton.

 

 

ON THE SURFACE, A HUMAN BEING and a chimpanzee might not seem to have much in common. However, they share 98% of their DNA. Chimpanzees are humans’ closest living relatives, and both primates often have similar facial expres- sions, emotions, and body movements. In these and many other ways, these two primates share a considerable amount of biology.

 

 

43 43

3 What is the genetic code?

What does the genetic code (DNA) do?

What is the genetic basis for human variation?

Genetics Reproducing Life and Producing Variation

There is a revolution going on in science: the discovery of DNA and the identifi-cation of its molecular structure have brought about a “DNA Revolution.” At no time in history have humans learned so much so quickly about the biology of plants and animals. In addition to bringing about developments in agriculture and food production, medicine, and other areas that affect billions of people every day, the infor- mation derived from DNA has transformed a number of scientific disciplines. Consider forensic science, where fingerprints and blood types were once the primary evidence. Thanks to DNA, far smaller samples— of tissue, bone, hair, and blood— can be used to identify victims’ remains and to identify criminals with far greater accuracy. DNA in samples saved from old crime scenes has helped free scores of individuals con- victed of crimes they had not committed. Beyond forensics, DNA analysis has helped determine family relationships. It has helped genealogists reach into the past to chart ancestry. It has even been used to detect the presence of diseases, such as leprosy and syphilis, in ancient skeletons. Given the long and growing list of ways in which DNA can be used, no wonder former US president Bill Clinton referred to the human DNA sequence, right after it was presented to the public in 2003, as “the most important, most wondrous map ever produced by mankind.”

When I studied introductory biology in college, in the early 1970s, knowledge of DNA was just a tiny fraction of what it is today. Evolution was understood in terms of entire organisms and their biological history. Now, DNA provides us with the information— a whole new window— whereby we can see how organisms are put together and what is actually evolving. Powerful stuff! In anthropology, it has meant new insights into

1

2

3

B I G Q U E S T I O N S ?

 

 

44 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

primate and human evolution. Before we can tie together the growing strands of DNA and evolution, though, we need to back up and examine the foundational work in genetics— the study of heredity.

Although the great nineteenth- century biologists discussed in chapter 2 knew a lot about variation in species, they did not fully understand how this variation is produced or how it is transmitted from parents to offspring. For example, how do an organism’s attributes grow from a fertilized egg? The answers to questions about variation— its ori- gin and continuation— lie in the cell, its structures, and the myriad functions it performs from conception through full maturity. And governing each cell is the genetic code.

The Cell: Its Role in Reproducing Life and Producing Variation The cell is the basic unit of life for all organisms (Figure  3.1). Every organism has at least one cell (that is the baseline definition of an organism). Organisms having cells with no internal compartments are called prokaryotes. These were likely the first life on Earth, appearing about 3.5 billion years ago (bya). Today, the prokaryotes are single- cell bacteria. Organisms with internal compartments sepa- rated by membranes are called eukaryotes. The membranes enclose the two main

prokaryotes Single- celled organisms with no nuclear membranes or organelles and with their genetic material as a single strand in the cytoplasm.

eukaryotes Multicelled organisms that have a membrane- bound nucleus con- taining both the genetic material and specialized organelles.

The nucleus is the largest organelle in a cell. It houses one copy of nearly all the genetic material, or DNA, of that organism. It is covered by a nuclear membrane, or nuclear envelope, which keeps the contents of the nucleus separate from the rest of the cell. The cell membrane is a semipermeable

membrane surrounding the entire cell, separating one cell from the next.

The mitochondrion is considered the “powerhouse” of the cell, because it generates most of the energy. The number of mitochondria per cell varies by tissue type and by organism.

The endoplasmic reticulum is an organelle that usually surrounds the nucleus. It plays an especially important role in protein synthesis (a process discussed later in this chapter).

The cytoplasm is fluid that fills the cell and maintains the cell’s shape. Organelles are suspended in the cytoplasm, which can also store chemical substances. The extranuclear DNA is in the mitochondria.

FIGURE 3.1 Cells and Their Organelles This illustration depicts the many components of cells found in plants and animals. Among the components are organelles, specialized parts analogous to organs.

 

 

The Cell: Its Role in Reproducing Life and Producing Variation | 45

parts of individual cells, the nucleus and the cytoplasm, between which various communications and activities happen (Figure  3.2). Eukaryotes evolved much later than prokaryotes, appearing some 1.2 bya. Their quite complex structures require enormous amounts of energy to survive and reproduce. As they did in the past, eukaryotes come in many different forms, ranging from single- cell yeasts to large, complex, multicellular organisms, such as us.

nucleus A membrane- bound structure in eukaryotic cells that contains the genetic material.

cytoplasm The jellylike substance inside the cell membrane that surrounds the nucleus and in which the organelles are suspended.

FIGURE 3.2 Prokaryotes and Eukaryotes (a) The many types of bacteria that we encounter in our daily lives are prokaryotic cells like this one. (b) For example, Escherichia coli (E. coli), two single cells of which are shown here, is a bacterium that aids digestion in the intestines of mammals, including humans. (c) This image shows the eukaryotic cells of a primate’s kidney.

Outer membrane

FlagellaeRibosome

Fimbriae

Cell wall Plasma membrane

Cytoplasm

The nucleoid region houses the genetic material of the prokaryotic cell, but unlike the nucleus of a eukaryotic cell it is not contained within a membrane. A prokaryotic cell has about one- thousandth the genetic material of a eukaryotic cell.

The cell wall provides a rigid shape and controls the movement of molecules into and out of the cell.

The flagellum is a whiplike structure attached to some prokaryotes. Rotated by a motorlike system located in the outer layers of the cell, the flagellum enables locomotion.

(a)

Nucleus Cytoplasm Plasma membrane (also called cell membrane)

(c)(b)

 

 

46 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

In all animals and plants, there are two types of eukaryotic cells. Somatic cells, also called body cells, comprise most tissues, such as bone, muscle, skin, brain, lung, fat, and hair (Figure 3.3). Gametes are the sex cells, sperm in males and ova (or eggs) in females (Figure  3.4). The root of somatic cell and gamete production is in the chromosomes, located in the nucleus of each cell. In humans, somatic cells have 46 chromosomes and gametes have 23 (Figure 3.5).

The DNA Molecule: The Genetic Code The chemical that makes up each chromosome, DNA, is the body’s genetic code. Because chromosomal DNA is contained in the nucleus of the cell, it is referred to as nuclear DNA, or nDNA.  Within each chromosome, DNA molecules form a sequence, or code, that is a template for the production of a protein, or part of a protein, in the body. Each protein has a specific function, and collectively the proteins determine all physical characteristics and govern the functions of all cells, tissues, and organs. Each DNA sequence, each protein- generating code, is a gene; and the complete set of genes in an individual cell is called the genome.

Although the number of chromosomes varies according to species (see Figure  3.5), all organisms share much the same genome. Chimpanzees have two more chromosomes than humans, but the DNA in chimpanzees and in humans is about 98% identical. Even the DNA in baker’s yeast is 45% similar to human  DNA.  Within any organism, nDNA is homoplasmic, meaning it is the same in each and every cell— the DNA in a skin cell matches the DNA in a bone cell. (An exception to the rule is mature red blood cells, which have no nuclei and, hence, no nuclear DNA.)

FIGURE 3.3 Somatic Cells Somatic cells in different tissues have different characteristics, but most somatic cells share a number of features. Every somatic cell has a nucleus, which contains a complete copy of the organism’s DNA. As a result, throughout the organism’s body there are millions of copies of that DNA. Note the nuclei in these images of human anatomy: (a) brain tissue, (b) red blood cells (the larger cells are white blood cells, and the small dots are platelets), (c) osteocyte (bone cell), (d) skin cells.

(a) (b)

(c) (d)

FIGURE 3.4 Gametes Only one of the sperm surrounding this ovum will penetrate the external membrane and fertilize the ovum.

somatic cells Diploid cells that form the organs, tissues, and other parts of an organism’s body.

gametes Sexual reproductive cells, ova and sperm, that have a haploid number of chromosomes and that can unite with a gamete of the opposite type to form a new organism.

homoplasmic Refers to nuclear DNA, which is identical in the nucleus of each cell type (except red blood cells).

 

 

The DNA Molecule: The Genetic Code | 47

A small but significant amount of DNA is contained in tiny organelles, called mitochondria, within each cell’s cytoplasm. These structures use oxygen to turn food molecules, especially sugar and fat, into adenosine triphosphate (ATP), a high- energy molecule that powers cells and, in turn, powers every tissue in the body. The number of mitochondria in a cell varies according to the cell’s

Organism Chromosome

Number Organism Chromosome

Number

FIGURE 3.5 Chromosomes (a) To get an idea of the incredibly minute size of chromosomes, consider that this pair has been magnified 35,000 times. If a penny (approximately 2 cm, or .8 in, in diameter) were magnified 35,000 times, it would be approximately .7 km, or .44 mi, in diameter. (b) An organism’s complexity is not related to its number of chromosomes, as this comparison illustrates. While humans have 46 chromosomes, other primates have more (e.g., ring- tailed lemurs) or fewer (e.g., black- and- white colobus monkeys).

Camel: 70

Guinea pig: 64

Salamander: 24

Housefly: 12

Apple: 34

Potato: 48

Petunia: 14

Algae: 148

Ring-tailed lemur: 56

Black-and- white colobus monkey: 44

Orangutan: 48

(b)

(a)

Camel: 70

Guinea pig: 64

Salamander: 24

Housefly: 12

Apple: 34

Potato: 48

Petunia: 14

Algae: 148

Ring-tailed lemur: 56

Black-and- white colobus monkey: 44

Orangutan: 48

(b)

(a)

mitochondria Energy- producing (ATP) organelles in eukaryotic cells; they pos- sess their own independent DNA.

adenosine triphosphate (ATP) An import- ant cellular molecule, created by the mitochondria and carrying the energy necessary for cellular functions.

 

 

48 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

activity level. For example, the cells in highly active body tissues, such as muscles, contain far more mitochondria than do cells in relatively inactive tissues, such as hair.

The mitochondrial DNA (mtDNA), a kind of miniature chromosome contain- ing 37 genes, is inherited just from the mother. That is, the mtDNA comes from the ovum. Each of us, then, carries our mother’s mtDNA, she carries or carried her mother’s mtDNA, and so on for generation after generation. In theory, a maternal lineage, or matriline, can be traced back hundreds of thousands of years. (Ancient matrilines in fossil hominins are discussed in chapter 10.) Unlike nDNA, mtDNA is heteroplasmic, meaning it can differ among different parts of a person’s body or even within the same kinds of cells.

DNA: THE BLUEPRINT OF LIFE The DNA molecule is the blueprint of life. It serves as the chemical template for every aspect of biological organisms. As Watson and Crick discovered, the molecule has a right- twisted, double- helix structure (see “DNA: Discovery of the Molecular Basis of Evolution” in chapter  2). Understanding this structure is key to understanding the growth of any organism and the transmission of genes from parents to offspring. The starting point for looking at the DNA molecule in any detail is to unravel a chromosome and look at a tiny segment of it under supermagnification. Its helical, ladderlike structure consists of two uprights and many rungs. The uprights of the structure are made up of alternating sugar and phosphate molecules, while the rungs are composed of paired nitrogen bases linked by a weak hydrogen bond (Figure 3.6).

On each side of the ladder, every unit of sugar, phosphate, and nitrogen base forms a single nucleotide (Figure  3.7). While the sugar and phosphate are the same throughout DNA, the base can be adenine (A), thymine (T), guanine (G), or cytosine (C). Owing to the bases’ unvarying chemical configurations, adenine and thymine always pair with each other and guanine and cytosine always pair up. In other words, apart from the rare errors in matching, adenine and thymine are complementary bases and guanine and cytosine are complementary bases. This means that if on one side of the ladder the sequence is ATGCAG, on the other side the sequence will be complementary, TACGTC. This predictability of base pair- ings assures the high reliability of one key function of the DNA molecule, that of self- reproduction. Anthropologists and geneticists investigate the many thousands of single DNA base pairings that produce genetic differences between individuals. Known as single nucleotide polymorphisms (SNPs, pronounced “snips”), these pairings are spread uniformly throughout the genome. Groups of SNPs play a crit- ical role in determining various attributes, such as hair color and blood type (see “Polymorphisms: Variations in Specific Genes,” later in this chapter).

The DNA Molecule: Replicating the Code One function of the DNA molecule is to replicate itself. Replication takes place in the nucleus and is part of cell division, leading to the production of new somatic cells (mitosis) or the production of new gametes (meiosis). Replication thus results

matriline DNA, such as mitochondrial DNA, whose inheritance can be traced from mother to daughter or to son.

heteroplasmic Refers to a mixture of more than one type of organellar DNA, such as mitochondrial DNA, within a cell or a single organism’s body, usually due to the mutation of the DNA in some organelles but not in others.

FIGURE 3.6 The Structure of DNA

The compact chromo- somal packaging of DNA enables thousands of genes to be housed inside a cell’s nucleus. The unwinding of this packaging reveals the genetic material.

DNA includes four different types of nitrogen bases. A gene is a specific and unique sequence of these bases.

DNA includes only one type of sugar (deoxyribose, the first part of the chemical name of DNA) and one type of phosphate group.

A T

GC

G C

P

PP

P

P

S

S

S S

S

S

 

 

The DNA Molecule: Replicating the Code | 49

in continued cell production, from the single- celled zygote (the fertilized egg) to two cells, then four cells, and so on, to the fully mature body with all of its many different tissues and organs— within which cells are continuously dying and being replaced.

In replication, DNA makes identical copies of itself, going from one double- stranded parent molecule of DNA to two double strands of daughter DNA. This means that where there was one chromosome, now there are two (Figure 3.8).

Nitrogen baseSugar

Phosphate

FIGURE 3.7 Nucleotide A nucleotide is the building block of DNA, made up of a phosphate group, a sugar, and a single nitrogen base.

complementary bases The predictable pairing of nitrogen bases in the structure of DNA and RNA, such that adenine and thymine always pair together (adenine and uracil in RNA) and cytosine and guanine pair together.

single nucleotide polymorphisms (SNPs) Variations in the DNA sequence due to the change of a single nitrogen base.

replication The process of copying nuclear DNA prior to cell division so that each new daughter cell receives a com- plete complement of DNA.

mitosis The process of cellular and nuclear division that creates two identical diploid daughter cells.

meiosis The production of gametes through one DNA replication and two cell (and nuclear) divisions, creating four haploid gametic cells.

zygote The cell that results from a sperm’s fertilization of an ovum.

C T TA

G A AT

C T TA

G A AT

C T TA

G A AT

C T

T A

C T

T A

G A

A T

G A

A TC

T T

A

G A

A T

Old strand

New strand

Old strand

New strand

The two strands of DNA become the parent template strands for replication. Each strand will replicate, using its nitrogen bases to synthesize a complementary strand.

Replication begins with the separation of the two strands of DNA. Enzymes break the relatively weak hydrogen bonds that hold together the paired nitrogen bases. In effect, the DNA is “unzipped,” creating the two parent template strands.

Each parent strand serves as a template for the creation of a new complementary DNA strand. The exposed, unpaired nitrogen bases on the parent strands attract complementary free-floating nucleotides. The nitrogen bases of these nucleotides form hydrogen bonds with the existing nitrogen bases—for example, a free-floating nucleotide with a cytosine base will attach itself to a guanine base.

New strands

New strands forming

When all the nitrogen bases of the parent strands are paired with (formerly free-floating) nucleotides, replication is complete. There are now two complete DNA molecules, each consisting of one parent strand and one new strand.

FIGURE 3.8 The Steps of DNA Replication

 

 

 

Imagine if we could look at ancient organisms’ DNA to understand their evolution. In fact, new tech- nology is enabling anthropologi- cal geneticists to routinely extract DNA from the tissues (mostly bones and teeth) of ancient remains. This emerging field, called paleoge- netics, has been made possible by the development of polymerase chain reaction (PCR), a method of amplifying a tiny sequence of DNA for study by incrementally increas- ing the sizes of a billion copies made from a single template of DNA. PCR has opened new windows onto the genetics of ancient populations, including the identification of sex chromosomes, the documentation of diseases, and the isolation of unique repetitions of DNA segments. It has yielded insight into the genetic dis- similarity of Neandertals and modern humans, and it has enabled explora- tion into population origins and move- ments (both subjects are among the topics of chapter 12).

Anthropologists have long specu- lated about the origins of Native Ameri- cans (where they came from is another subject of chapter 12). Key to under- standing their origins is their genetic diversity. Studies have revealed that the haplogroups of mtDNA— A, B, C, and D— in living Native Americans are quite similar to the haplogroups of their ancestors. This resemblance strongly suggests that Native Amer- icans’ genetic structure is quite

By examining the distribution of haplogroups A, B, C, and D in North and South America as well as eastern Asia, researchers have estimated that Native Americans arrived in the West- ern Hemisphere between 15,000 and 40,000 yBP. The presence of the same haplogroups in the northeastern part of Asia suggests that Native Americans originated from this area.

A B

CD A

AB

D C AB

CD

Ancient DNA Opens New Windows on the Past

old. Based on current assumptions about mutation rates in mtDNA, anthropologists and geneticists have estimated that people arrived in the Americas sometime between 15,000 and 40,000 yBP, earlier than what is documented in the archaeo- logical record (among the subjects of chapter  13). But a new variant of haplogroup D, discovered in the DNA of a 10,300- year- old skeleton from Alaska by paleogeneticist Brian Kemp and his collaborators, sug- gests that the molecular clock may be off and that humans first arrived

in the New World around 13,500 yBP, a date that jibes well with the archaeological evidence.

In addition, the presence of all four haplogroups and their variants in the skeletons of Native Ameri- cans dating to before 1492 tells us that the widespread decline of the Native American population after Columbus’s arrival did not reduce their genetic, and therefore biologic, diversity. Native Americans living today are likely as diverse genetically as were their ancestors living hun- dreds and thousands of years ago.

H O W D O W E K N O W ?

 

 

CHROMOSOME TYPES Within somatic cells, chromosomes occur in homologous, or matching, pairs (Figure  3.9). Each pair includes the father’s contribution (the paternal chromo- some) and the mother’s contribution (the maternal chromosome). These nonsex chromosomes are called autosomes.

The karyotype, or complete set of chromosomes, includes all of the autosomes and one pair of sex chromosomes, so called because they determine an individ- ual’s biological sex. Females have two X chromosomes, and males have one X and one Y chromosome (Figure 3.10). The Y chromosome contains a small amount of genetic material, which determines only male characteristics. The interaction of

FIGURE 3.9 Chromosome Pairs Homologous chromosomes are virtually identical in their physical and chemical structure. Each pair of chromosomes has the same genes, but the pair may have different alleles for specific genes.

FIGURE 3.10 Karyotype Contained within each somatic cell, the human karyotype typically consists of 46 chromosomes of various sizes in 23 pairs. Of those 23, one pair determines the person’s sex. Here, the label “X” means that these sex chromosomes are both Xs and thus belong to a human female.

free- floating nucleotides Nucleotides (the basic building block of DNA and RNA) that are present in the nucleus and are used during DNA replication and mRNA synthesis.

homologous Refers to each set of paired chromosomes in the genome.

autosomes All chromosomes, except the sex chromosomes, that occur in pairs in all somatic cells (not the gametes).

karyotype The characteristics of the chro- mosomes for an individual organism or a species, such as number, size, and type. The karyotype is typically presented as a photograph of a person’s chromosomes that have been arranged in homologous pairs and put in numerical order by size.

sex chromosomes The pair of chromo- somes that determine an organism’s biological sex.

The Two Steps of DNA Replication

The first function of DNA is to replicate itself. Templates from the original (parental) strand of DNA yield two exact (daughter) copies.

Step Activity

1. Strand of DNA unzips to form templates.

Weak nucleotide bonds between bases break, exposing two parental strands of DNA.

2. Templates plus nucleotides yield daughters.

Free- floating nucleotides in the nucleus match with the newly exposed template strands of DNA.

C O N C E P T C H E C K !

The DNA Molecule: Replicating the Code | 51

 

 

52 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

gametes during fertilization determines the combination of chromosomes in the offspring. If an X- carrying sperm fertilizes an egg (which always carries an X), the offspring will be female. If a Y- carrying sperm fertilizes an egg, the offspring will be male. Therefore, the male parent’s gamete determines the sex of his offspring because the Y chromosome is present in males only; it is passed from father to son.

The Y chromosome can be highly informative about paternity over many gen- erations. This patriline is in part analogous to the matriline- based mtDNA, which is passed on only by females. However, mtDNA goes to all of a woman’s children, whereas the Y chromosome is passed only to a man’s son.

Mitosis: Production of Identical Somatic Cells An organism starts life as a single cell, the zygote, which then produces identical copies of itself many, many times. A single human zygote, for example, even- tually results in more than 10 trillion cells, each having the exact same DNA (Figure  3.11). Here, the production of identical daughter cells from an original

patriline DNA whose inheritance can be traced from father to son via the Y chromosome.

Fertilization occurs when one sperm penetrates the outer membrane of the ovum, or egg.

A zygote is the single cell that forms at the beginning of an organism’s life. This cell must replicate itself millions of times to form a fully developed fetus.

After the regions of the body are established and different types of tissue have formed, the fetus grows until it reaches full-term size.

Following many replications, the embryo begins to differentiate different types of tissues and separate regions of the body, such as the head and limbs.

FIGURE 3.11 Prenatal Development The stages of human development from fertilization to a full- term infant.

diploid cell A cell that has a full comple- ment of paired chromosomes.

 

 

parental cell, mitosis, involves one DNA replication followed by one cell division (Figure 3.12). In this kind of cell division, a diploid cell— a cell having its organ- ism’s full set of chromosomes— divides to produce two cells, each of which also has the full set of chromosomes.

Chromatin

46 chromosomes

46 chromosomes

46 chromosomes

Chromosome duplication

Line up, spindle forms

The DNA in the cell is replicated, as described in Figure 3.8. This parent cell has a full set (23 pairs) of chromosomes.

The replicated DNA lines up in the middle of the cell.

As the cell divides, the DNA sep- arates. One complete complement of DNA goes into one new cell, and the other full complement goes into the second new cell.

Each new cell has a full set of DNA, with 23 pairs of chromosomes.

(b)

(a)

FIGURE 3.12 Mitosis (a) The steps of mitosis in humans. (b) A human skin cell undergoing mitosis, dividing into two new daughter cells.

46 chromo- somes

Chromo- some

duplication

Homologous chromosome

pairing

The pairs are separated and pulled to opposite sides as the cell divides.

The cell divides, forming two daughter cells.

The chromosomes line up in the middle of the cell, but are not paired.

A second cell division follows, without DNA replication. The chromosomes are separated into single strands and are pulled to opposite sides as the cell divides.

Four haploid daughter cells result, each with 23 chromosomes, but no pairs.

The parent cell is diploid, hav- ing 23 pairs of chromosomes.

Meiosis begins like mitosis. Initially, the cell’s DNA is replicated.

Homologous pairs of chromosomes line up in the middle of the cell.

FIGURE 3.13 Meiosis The steps of meiosis in humans. Compare this process to mitosis, shown in Figure 3.12.

Mitosis: Production of Identical Somatic Cells | 53

 

 

54 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

Meiosis: Production of Gametes (Sex Cells) The genetic code is transmitted from parents to offspring via the female and male gametes. Gametes, remember, have only half the chromosomes that are in somatic cells— they are haploid, containing one chromosome from each pair. Unlike mito- sis, the production of these cells, meiosis, does not result in identical copies of the parent cell and the parent cell’s DNA. Meiosis involves one DNA replication followed by two cell divisions (Figure 3.13).

Meiosis plays a critical role in the inheritance of biological characteristics and the variation seen in offspring. Because each gamete contains just one chromosome from a homologous pair and just one sex chromosome, during reproduction each parent contributes only half of his or her genetic material. For example, in your somatic cells, each homologous pair includes one chromosome from your mother and one chromosome from your father. Whether a particular gamete contains your mother’s chromosome or your father’s chromosome is completely random. In addition, homologous chromosomes often exchange parts when they pair up and intertwine. This exchange of parts is called crossing- over. The outcome of such reshuffling is that gene variants originally on the maternal chromosome are now on the paternal chromosome (or vice versa), a common development called recombination. Genes that are close together on a chromosome are much less likely to recombine. These units or blocks of genetic material are called haplo- types. Geneticists prefer to study haplotypes because they do not recombine and are passed on for many generations, potentially hundreds, over time. Groups of related haplotypes, called haplogroups, are an important tool for studying both long- term evolution and populations’ histories.

In rare instances, during meiosis nonhomologous chromosomes exchange seg- ments. These rare exchanges are called translocations. The most common form in humans, involving both chromosome 13 and chromosome 14, affects about 1 in 1,300 people. Translocations may cause infertility, Down syndrome (when one- third of chromosome 21 joins onto chromosome 14), and a number of diseases, including several forms of cancer (some leukemias). On occasion, chromosome pairs fail to separate during meiosis or mitosis. These nondisjunctions result in an incorrect number of chromosomes in the person’s genome. A loss in number of chromosomes is a monosomy. A gain in number of chromosomes is a trisomy, the most common being trisomy- 21, or Down syndrome (in this form, caused by an extra or part of an extra chromosome 21). As with many chromosomal abnor- malities, the age of the mother determines the risk of the offspring’s having Down syndrome. For 20- to 24- year- old mothers, the risk is 1/1,490. It rises to 1/106 by age 40 and 1/11 beyond age 49.

As Mendel had recognized (see “Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA” in chapter  2), each physical unit (that is, gene) passes from parent to offspring independently of other physical units. This independent inheritance— often called Mendel’s law of independent assortment (Figure  3.14)—applies to genes from different chromosomes. How- ever, what happens when genes are on the same chromosome? Because meiosis involves the separation of chromosome pairs (homologous chromosomes), genes on the same chromosome, especially ones near each other on that chromosome, have a greater chance of being inherited as a package. They are less subject to

haploid cell A cell that has a single set of unpaired chromosomes; half of the num- ber of chromosomes as a diploid cell.

crossing- over The process by which homologous chromosomes partially wrap around each other and exchange genetic information during meiosis.

recombination The exchange of genetic material between homologous chromo- somes, resulting from a cross- over event.

haplotypes A group of alleles that tend to be inherited as a unit due to their closely spaced loci on a single chromosome.

haplogroups A large set of haplotypes, such as the Y chromosome or mitochon- drial DNA, that may be used to define a population.

translocations Rearrangements of chromosomes due to the insertion of genetic material from one chromosome to another.

nondisjunctions Refers to the failure of the chromosomes to properly segregate during meiosis, creating some gametes with abnormal numbers of chromosomes.

monosomy Refers to the condition in which only one of a specific pair of chro- mosomes is present in a cell’s nucleus.

trisomy Refers to the condition in which an additional chromosome exists with the homologous pair.

law of independent assortment Mendel’s second law, which asserts that the inher- itance of one trait does not affect the inheritance of other traits.

F1 GgYy

F1 GgYy

GY

GY

Gy

gY

gy

GGYy GGyy

Gy gY gy

GGYY GGYy

GgYy Ggyy

GgYY GgYy

GgYy Ggyy

GgYY GgYy

ggYy ggyy

ggYY ggYy

This Punnett square shows all possible combinations of two different genes, pod color and seed color: G = green pod g = yellow pod Y = yellow seed y = green seed

Since both parents have two different alleles for both traits, there are 16 possible combinations.

Of the 16 combinations, nine have green pods and yellow seeds. The genotypes vary and include: GGYY, GGYy, GgYY, GgYy.

Three of the resulting combinations have green pods and green seeds. They have one of two genotypes: GGyy or Ggyy.

Three of the resulting combinations have yellow pods and yellow seeds. There are two possible genotypes: ggYY or ggYy.

One of the resulting combinations has yellow pods and green seeds, with the genotype ggyy.

FIGURE 3.14 Law of Independent Assortment Through his research with pea plants, Mendel created several laws pertaining to inheritance. (a) His second law, the law of independent assortment, asserts that traits linked to different chromosomes are inherited independently from one another. (b) Hair color, for example, is inherited independently from eye color.

Two equally probable chromosome arrangements in

meiosis I: OR

meiosis II: OR

gametes OR

with genotypes OR

The large chromosomes have a gene that determines eye color: red represents brown eyes, and blue represents blue eyes.

Alternatively, following the first cell division of meiosis, one cell has genes for brown eyes and blond hair, while the second has genes for blue eyes and brown hair.

Again, the four daughter cells, or gametes, follow suit: two gametes have genes for brown eyes and blond hair, while the other two gametes have genes for blue eyes and brown hair.

The small chromosomes have a gene that determines hair color: red represents brown hair, and blue represents blond hair.

Following the first cell division of meiosis, one cell has genes for brown eyes and brown hair, while the second cell has genes for blue eyes and blond hair.

The four daughter cells, or gametes, follow suit: two gametes have genes for brown eyes and brown hair, while the other two gametes have genes for blue eyes and blond hair.

(a)

(b)

 

 

Meiosis: Production of Gametes (Sex Cells) The genetic code is transmitted from parents to offspring via the female and male gametes. Gametes, remember, have only half the chromosomes that are in somatic cells— they are haploid, containing one chromosome from each pair. Unlike mito- sis, the production of these cells, meiosis, does not result in identical copies of the parent cell and the parent cell’s DNA. Meiosis involves one DNA replication followed by two cell divisions (Figure 3.13).

Meiosis plays a critical role in the inheritance of biological characteristics and the variation seen in offspring. Because each gamete contains just one chromosome from a homologous pair and just one sex chromosome, during reproduction each parent contributes only half of his or her genetic material. For example, in your somatic cells, each homologous pair includes one chromosome from your mother and one chromosome from your father. Whether a particular gamete contains your mother’s chromosome or your father’s chromosome is completely random. In addition, homologous chromosomes often exchange parts when they pair up and intertwine. This exchange of parts is called crossing- over. The outcome of such reshuffling is that gene variants originally on the maternal chromosome are now on the paternal chromosome (or vice versa), a common development called recombination. Genes that are close together on a chromosome are much less likely to recombine. These units or blocks of genetic material are called haplo- types. Geneticists prefer to study haplotypes because they do not recombine and are passed on for many generations, potentially hundreds, over time. Groups of related haplotypes, called haplogroups, are an important tool for studying both long- term evolution and populations’ histories.

In rare instances, during meiosis nonhomologous chromosomes exchange seg- ments. These rare exchanges are called translocations. The most common form in humans, involving both chromosome 13 and chromosome 14, affects about 1 in 1,300 people. Translocations may cause infertility, Down syndrome (when one- third of chromosome 21 joins onto chromosome 14), and a number of diseases, including several forms of cancer (some leukemias). On occasion, chromosome pairs fail to separate during meiosis or mitosis. These nondisjunctions result in an incorrect number of chromosomes in the person’s genome. A loss in number of chromosomes is a monosomy. A gain in number of chromosomes is a trisomy, the most common being trisomy- 21, or Down syndrome (in this form, caused by an extra or part of an extra chromosome 21). As with many chromosomal abnor- malities, the age of the mother determines the risk of the offspring’s having Down syndrome. For 20- to 24- year- old mothers, the risk is 1/1,490. It rises to 1/106 by age 40 and 1/11 beyond age 49.

As Mendel had recognized (see “Since Darwin: Mechanisms of Inheritance, the Evolutionary Synthesis, and the Discovery of DNA” in chapter  2), each physical unit (that is, gene) passes from parent to offspring independently of other physical units. This independent inheritance— often called Mendel’s law of independent assortment (Figure  3.14)—applies to genes from different chromosomes. How- ever, what happens when genes are on the same chromosome? Because meiosis involves the separation of chromosome pairs (homologous chromosomes), genes on the same chromosome, especially ones near each other on that chromosome, have a greater chance of being inherited as a package. They are less subject to

haploid cell A cell that has a single set of unpaired chromosomes; half of the num- ber of chromosomes as a diploid cell.

crossing- over The process by which homologous chromosomes partially wrap around each other and exchange genetic information during meiosis.

recombination The exchange of genetic material between homologous chromo- somes, resulting from a cross- over event.

haplotypes A group of alleles that tend to be inherited as a unit due to their closely spaced loci on a single chromosome.

haplogroups A large set of haplotypes, such as the Y chromosome or mitochon- drial DNA, that may be used to define a population.

translocations Rearrangements of chromosomes due to the insertion of genetic material from one chromosome to another.

nondisjunctions Refers to the failure of the chromosomes to properly segregate during meiosis, creating some gametes with abnormal numbers of chromosomes.

monosomy Refers to the condition in which only one of a specific pair of chro- mosomes is present in a cell’s nucleus.

trisomy Refers to the condition in which an additional chromosome exists with the homologous pair.

law of independent assortment Mendel’s second law, which asserts that the inher- itance of one trait does not affect the inheritance of other traits.

F1 GgYy

F1 GgYy

GY

GY

Gy

gY

gy

GGYy GGyy

Gy gY gy

GGYY GGYy

GgYy Ggyy

GgYY GgYy

GgYy Ggyy

GgYY GgYy

ggYy ggyy

ggYY ggYy

This Punnett square shows all possible combinations of two different genes, pod color and seed color: G = green pod g = yellow pod Y = yellow seed y = green seed

Since both parents have two different alleles for both traits, there are 16 possible combinations.

Of the 16 combinations, nine have green pods and yellow seeds. The genotypes vary and include: GGYY, GGYy, GgYY, GgYy.

Three of the resulting combinations have green pods and green seeds. They have one of two genotypes: GGyy or Ggyy.

Three of the resulting combinations have yellow pods and yellow seeds. There are two possible genotypes: ggYY or ggYy.

One of the resulting combinations has yellow pods and green seeds, with the genotype ggyy.

FIGURE 3.14 Law of Independent Assortment Through his research with pea plants, Mendel created several laws pertaining to inheritance. (a) His second law, the law of independent assortment, asserts that traits linked to different chromosomes are inherited independently from one another. (b) Hair color, for example, is inherited independently from eye color.

Two equally probable chromosome arrangements in

meiosis I: OR

meiosis II: OR

gametes OR

with genotypes OR

The large chromosomes have a gene that determines eye color: red represents brown eyes, and blue represents blue eyes.

Alternatively, following the first cell division of meiosis, one cell has genes for brown eyes and blond hair, while the second has genes for blue eyes and brown hair.

Again, the four daughter cells, or gametes, follow suit: two gametes have genes for brown eyes and blond hair, while the other two gametes have genes for blue eyes and brown hair.

The small chromosomes have a gene that determines hair color: red represents brown hair, and blue represents blond hair.

Following the first cell division of meiosis, one cell has genes for brown eyes and brown hair, while the second cell has genes for blue eyes and blond hair.

The four daughter cells, or gametes, follow suit: two gametes have genes for brown eyes and brown hair, while the other two gametes have genes for blue eyes and blond hair.

(a)

(b)

Meiosis: Production of Gametes (Sex Cells) | 55

 

 

56 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

recombination. This gene linkage— the inheritance of a package of genes (such as haplotypes) from the same chromosome (Figure 3.15)—is an exception to Mendel’s law of independent assortment.

Producing Proteins: The Other Function of DNA In addition to replicating itself, DNA serves as the template for protein synthe- sis. Proteins are the complex chemicals that make up tissues and bring about the functions, repair, and growth of tissues (Table  3.1). While some work within cells— for example, the enzymes that unzip DNA during replication— others, such as hormones, work within the whole body. Proteins consist of amino acids, of which there are 20. Each kind of protein is defined by its particular combination and number of linked amino acids. Most of the human body is comprised of pro- teins, and the body produces 12 of the amino acids. The other eight, also called essential amino acids, come from particular foods.

Two main categories of proteins are constantly being synthesized. Structural proteins are responsible for physical characteristics, such as hair form, eye color, tooth size, and basic bone shape (Figure  3.16). The other category, regulatory (also called functional ) proteins, includes enzymes, hormones, and antibodies. Enzymes regulate activities within cells, hormones regulate activities between cells, and antibodies are key to fighting infections.

Protein synthesis is a two- step process (Figure 3.17). The first step, transcrip- tion, takes place mostly in the cell’s nucleus. The second, translation, takes place in the cytoplasm. Transcription starts out just like the first step of DNA replica- tion: a double strand of parental DNA unzips. Rather than producing daughter strands of DNA, the now- exposed bases in the DNA molecule serve as a single

A B C

a b c

FIGURE 3.15 Linkage Since alleles A, B, and C are on the same chromosome, they have a better chance of being inherited as a unit than of being combined and inherited with alleles a, b, and c, which are together on a separate chromosome and thus also have a good chance of being inherited as a unit. Meanwhile, because they are close together, alleles A and B (like alleles a and b) stand a better chance of being inherited together than do alleles B and C (like alleles b and c). If, for example, eye color and hair color were on the same chromosome, especially if they were close together on that chromosome, they would most likely not be inherited separately with all the combinations examined in Figure 3.14.

linkage Refers to the inheritance, as a unit, of individual genes closely located on a chromosome; an exception to the law of independent assortment.

amino acids Organic molecules combined in a specific sequence by the ribosomes to form a protein.

essential amino acids Those amino acids that cannot be synthesized in the body; they must be supplied by the diet.

structural proteins Proteins that form an organism’s physical attributes.

regulatory proteins Proteins involved in the expression of control genes.

transcription The first step of protein synthesis, involving the creation of mRNA based on the DNA template.

translation The second step of protein synthesis, involving the transfer of amino acids by tRNA to the ribosomes, which are then added to the protein chain.

ribonucleic acid (RNA) A single- stranded molecule involved in protein synthesis, consisting of a phosphate, ribose sugar, and one of four nitrogen bases.

TABLE 3.1 The Seven Types of Proteins

Name Function Examples

Enzymes Catalyze chemical reactions

Lactase— breaks down lactose in milk products

Structural Proteins

Give structure or support to tissues

Keratin— hair; collagen— bone

Gas Transport Proteins

Carry vital gases to tissues

Hemoglobin— oxygen

Antibodies Part of immune system Anti- A and anti- B in ABO blood system

Hormones Regulate metabolism Insulin— regulates metabolism of carbohydrates and fats

Mechanical Proteins

Carry out specific functions or work

Actin and myosin— help muscles contract

Nutrients Provide vital nutrients to tissues

Ovalbumin— main protein of egg whites

 

 

Producing Proteins: The Other Function of DNA | 57

template for another kind of nucleic acid, ribonucleic acid (RNA). RNA has the same nitrogen bases as DNA, except that uracil (U) replaces thymine (T). Uracil always matches with adenine (A), while guanine (G) continues to pair with cytosine (C).

Only one of the two DNA strands serves as the template for the production of RNA. This strand attracts free- floating RNA nucleotides. The strand of RNA— now called messenger RNA (mRNA)—then splits off from the DNA template, leaves the nucleus, and moves into the cytoplasm.

In the translation step, the mRNA attaches itself to structures called ribo- somes. The mRNA is a “messenger” because (in the form of its own open bases) it carries the code for the protein being synthesized from the nucleus to the ribo- somes. Ribosomes are made up of another kind of ribonucleic acid, ribosomal RNA (rRNA). Once the mRNA is attached to the ribosome, the transcription step of protein synthesis is complete.

Floating in the cytoplasm is yet another kind of ribonucleic acid, transfer RNA (tRNA). tRNA occurs as triplets, or anticodons, that seek complementary triplet strands of mRNA, known as triplets or codons. For example, a triplet of AUC mRNA would pair with the complementary UAG tRNA. The three bases of the tRNA triplet represent a specific amino acid.

As the tRNA strand builds off the mRNA template, the amino acids are chemically linked together by a peptide bond, resulting in a chain of amino acids. A chain of these peptide bonds is called a polypeptide. Although a single polypeptide may function as a protein, in many cases multiple polypeptides must bind together and fold into a three- dimensional structure to form a functional protein. For example, hemoglobin, a molecule found on the surface of red blood cells, is comprised of two pairs of polypeptide chains. Once the protein has formed, it breaks away from the tRNA and commences with its task, either structural or functional.

(a) (b)

FIGURE 3.16 Structural Proteins While culture and environment strongly influence the development of biological structures, those structures are initially determined by structural proteins. Two important structural proteins are keratin and collagen. (a) In humans, keratin is the primary component of hair (pictured here), skin, and fingernails. In other mammals and in amphibians, birds, and reptiles, it also contributes to structures such as hooves, antlers, claws, beaks, scales, and shells. (b) Collagen is the most abundant protein in humans and other mammals and is essential for connective tissues, such as bone (pictured here), cartilage, ligaments, and tendons. In addition, collagen strengthens the walls of blood vessels and, along with keratin, gives strength and elasticity to skin. In its crystalline form, collagen is found in the cornea and lens of the eye.

uracil One of four nitrogen bases that make up RNA; it pairs with adenine.

messenger RNA (mRNA) The molecules that are responsible for making a chemical copy of a gene needed for a specific pro- tein, that is, for the transcription phase of protein synthesis.

ribosomes The organelles attached to the surface of endoplasmic reticulum, located in the cytoplasm of a cell; they are the site of protein synthesis.

ribosomal RNA (rRNA) A fundamental structural component of a ribosome.

transfer RNA (tRNA) The molecules that are responsible for transporting amino acids to the ribosomes during protein synthesis.

anticodons Sequences of three nitrogen bases carried by tRNA, they match up with the complementary mRNA codons and each designate a specific amino acid during protein synthesis.

triplets Sequences of three nitrogen bases each in DNA, known as codons in mRNA.

codons The sequences of three nitrogen bases carried by mRNA that are coded to produce specific amino acids in protein synthesis.

peptide bond Chemical bond that joins amino acids into a protein chain.

polypeptide Also known as a protein, a chain of amino acids held together by multiple peptide bonds.

 

 

Amino acid

Protein

RibosomeNucleus

DNA

mRNA

Cytoplasm

Transcription, which occurs in the nucleus, involves the creation of mRNA from one strand of DNA.

After the mRNA strand is completed, it leaves the nucleus and goes to the ribosomes, in the cytoplasm.

Translation takes place at the ribo- somes. A protein is formed as the mRNA is “read” and the appropriate amino acids are linked together.

Transcription

Translation

Transcription Translation

As in DNA replication, transcription begins with enzymes “unzipping” the DNA. Unlike replication, however, transcription uses only one strand of DNA.

Once the DNA strands have opened, messenger RNA (mRNA) attaches free-floating RNA nitrogen bases to the exposed, unpaired DNA nitrogen bases.

Once completed, the DNA closes back up, and the mRNA strand leaves the nucleus and goes to one of the ribosomes on the endoplasmic reticulum.

At a ribosome, a molecule of tRNA brings the anticodon for each codon on the mRNA. The tRNA carries its anticodon on one end and the associated amino acid on the other.

Translation begins as the mRNA binds to a ribosome. In effect, the “message” carried by the mRNA is “translated” by a ribosome.

The ribosome “reads” the mRNA three nitrogen bases at a time. When a codon matches the transfer RNA (tRNA) molecule’s anticodon, the tRNA’s amino acid is added to the protein chain. For example, if the codon has the bases AUG, then the tRNA with the anticodon of UAC will attach the amino acid methionine to the chain.

mRNA

tRNA

(a)

(b)

DNA template

in nucleus

Ribosome

at ribosome

mRNA strand

A

U

T

A

T

A

C

G

G

C …

Completed mRNA strand

Moves out of nucleus to ribosomes in cytoplasm

U UUAA G AG A GG G ……

U

UU

A

A

G

A

G

A GG G ……

Anticodon

Codon

tRNA

Amino acid

AU C

… U

As the ribosome moves the mRNA one codon at a time, tRNA continues to attach the appropriate amino acid to the protein chain. The amino acids are attached by a peptide bond, creating a polypeptide chain, which when completed is the protein. As each amino acid is added, the tRNA is released.

Peptide bond

Protein: polypeptide chain

AU C A

U

C

UA G

UA G

Eventually, a “stop” codon is reached, which indicates that the protein is completed. The mRNA leaves the ribosome, and the protein is released.

A U C

Anticodon

Amino acid

Anticodon

Codon

(c)

met

glysermetleutyrlys

sermet

stop

Serine

UUA AG A GG G …

UU AA A GG G G …

AG G U A G …

UA G… Protein Synthesis (a) As this overview of protein synthesis shows, transcription occurs mostly within the cell’s nucleus and translation follows at the ribosomes. (b) Here, the steps of protein synthesis are diagrammed for a hypothetical protein. (c) At the top is a computer model of a tRNA molecule, and below that is a diagram of the molecule.

F I G U R E

3.17 Protein Synthesis

 

 

Amino acid

Protein

RibosomeNucleus

DNA

mRNA

Cytoplasm

Transcription, which occurs in the nucleus, involves the creation of mRNA from one strand of DNA.

After the mRNA strand is completed, it leaves the nucleus and goes to the ribosomes, in the cytoplasm.

Translation takes place at the ribo- somes. A protein is formed as the mRNA is “read” and the appropriate amino acids are linked together.

Transcription

Translation

Transcription Translation

As in DNA replication, transcription begins with enzymes “unzipping” the DNA. Unlike replication, however, transcription uses only one strand of DNA.

Once the DNA strands have opened, messenger RNA (mRNA) attaches free-floating RNA nitrogen bases to the exposed, unpaired DNA nitrogen bases.

Once completed, the DNA closes back up, and the mRNA strand leaves the nucleus and goes to one of the ribosomes on the endoplasmic reticulum.

At a ribosome, a molecule of tRNA brings the anticodon for each codon on the mRNA. The tRNA carries its anticodon on one end and the associated amino acid on the other.

Translation begins as the mRNA binds to a ribosome. In effect, the “message” carried by the mRNA is “translated” by a ribosome.

The ribosome “reads” the mRNA three nitrogen bases at a time. When a codon matches the transfer RNA (tRNA) molecule’s anticodon, the tRNA’s amino acid is added to the protein chain. For example, if the codon has the bases AUG, then the tRNA with the anticodon of UAC will attach the amino acid methionine to the chain.

mRNA

tRNA

(a)

(b)

DNA template

in nucleus

Ribosome

at ribosome

mRNA strand

A

U

T

A

T

A

C

G

G

C …

Completed mRNA strand

Moves out of nucleus to ribosomes in cytoplasm

U UUAA G AG A GG G ……

U

UU

A

A

G

A

G

A GG G ……

Anticodon

Codon

tRNA

Amino acid

AU C

… U

As the ribosome moves the mRNA one codon at a time, tRNA continues to attach the appropriate amino acid to the protein chain. The amino acids are attached by a peptide bond, creating a polypeptide chain, which when completed is the protein. As each amino acid is added, the tRNA is released.

Peptide bond

Protein: polypeptide chain

AU C A

U

C

UA G

UA G

Eventually, a “stop” codon is reached, which indicates that the protein is completed. The mRNA leaves the ribosome, and the protein is released.

A U C

Anticodon

Amino acid

Anticodon

Codon

(c)

met

glysermetleutyrlys

sermet

stop

Serine

UUA AG A GG G …

UU AA A GG G G …

AG G U A G …

UA G…

 

 

60 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

All of the DNA involved in protein synthesis is coding DNA, the molecular segments encoded for particular proteins. The total length of DNA in humans is about 3 billion nucleotides. Each of the 20,000 or so genes has about 5,000 nucle- otides, so (according to the math) only about 5% of the DNA contains protein- coding material. Thus, most human DNA is noncoding. Often interspersed with coding DNA from one end of the chromosome to the other, this noncoding “junk” DNA has long been thought to have no function. However, non- protein- coding DNA is now known to have considerable regulatory functions throughout the genome, specifically regulating gene activities by helping to turn them on or off. New work on the genome suggests that 80% of the noncoding DNA is functional in some manner, containing instructions for proteins such as which genes a cell uses and when or determining if a cell becomes a bone cell or a brain cell. In pro- tein synthesis, the noncoding DNA is cut out before translation. Recent studies by anthropological geneticists have suggested that noncoding DNA located close to genes that control brain function may have a role in the overall wiring of the brain cells to each other. Most research, however, focuses on the DNA that codes for particular body structures or particular regulatory functions. This DNA makes up the two main types of genes, structural and regulatory.

coding DNA Sequences of a gene’s DNA (also known as exons) that are coded to produce a specific protein and are transcribed and translated during protein synthesis.

noncoding DNA Sequences of a gene’s DNA (also known as introns) that are not coded to produce specific proteins and are excised before protein synthesis.

structural genes Genes coded to produce particular products, such as an enzyme or hormone, rather than for regulatory proteins.

regulatory genes Those genes that deter- mine when structural genes and other regulatory genes are turned on and off for protein synthesis.

homeotic (Hox) genes Also known as homeobox genes, they are responsible for differentiating the specific segments of the body, such as the head, tail, and limbs, during embryological development.

locus The location on a chromosome of a specific gene.

The Two Steps of Protein Synthesis

The second function of DNA is to synthesize proteins, which are responsible for all the structures and functions of the body.

Step Activity

1. Transcription (nucleus)

Parental strand of DNA unzips, exposing two daughter strands of DNA.

Free- floating RNA nucleotides match one exposed daughter strand of DNA.

The strand of messenger RNA (mRNA) moves out of the nucleus and into the cytoplasm.

2. Translation (cytoplasm)

The mRNA attaches to a ribosome in the cytoplasm.

Triplets of transfer RNA (tRNA), with exposed bases and each carrying an amino acid specific to its set of three bases, recognize and bind with complementary base pairs of mRNA.

The amino acids, linked by peptide bonds, form a chain called a polypeptide.

The protein forms, either as a single polypeptide or as multiple polypeptides bound together.

C O N C E P T C H E C K !

 

 

Polymorphisms: Variations in Specific Genes | 61

Genes: Structural and Regulatory Structural genes are responsible for body structures, such as hair, blood, and other tissues. Regulatory genes turn other genes on and off, an essential activity in growth and development. If the genes that determine bones, for example, did not turn off at a certain point, bones would continue to grow well beyond what would be acceptable for a normal life (Figure 3.18).

Chickens have the genes for tooth development, but they do not develop teeth because those genes are permanently turned off. Humans have a gene for complete body hair coverage, but that gene is not turned on completely. The human genes for sexual maturity turn on during puberty, somewhat earlier in girls than in boys. Finally, regulatory genes can lead to lactose intolerance in humans (among the topics of chapter 4). In this instance, the gene that produces lactase— the enzyme for the digestion of milk— is turned off for most human populations around the world following weaning, usually by about age four. However, most humans of northern European and East African descent have inherited a different regulatory gene, which creates lactase persistence. A person who lacks this gene and eats dairy products experiences great gastrointestinal discomfort. A person who retains the gene is able to digest lactose owing to the persistence of lactase, thus enjoying the nutritional benefits of milk.

An organism’s form and the arrangement of its tissues and organs are deter- mined by regulatory genes called homeotic (Hox) genes. These master genes guide, for example, the embryological development of all the regions of an animal’s body, such as the head, trunk, and limbs (Figure  3.19). This means that in the process of development particular sets of Hox genes are turned on in a particular sequence, causing the correct structure or part of a structure to develop in each region. Until recently, scientists thought that the genes that control the develop- ment of the key structures and functions of the body differed from organism to organism. We now know, however, that the development of various body parts in complex organisms— such as the limbs, eyes, and vital organs— is governed by the same genes. Hox genes were first found in fruit flies, but research has shown that a common ancestral lineage has given organisms— ranging from flies to mice to humans— the same basic DNA structure in the key areas that control the development of form. Flies look like flies, mice look like mice, and humans look like humans because the Hox genes are turned on and off at different places and different times during the development process.

Polymorphisms: Variations in Specific Genes Along each chromosome, a specific gene has a specific physical location, or locus (plural, loci). This locus is of intense interest to geneticists, especially in under- standing the appearance and evolution of genetic variation (among the topics of chapter 4). Alleles, the genetic subunits (see “Mechanisms of Inheritance” in chap- ter  2), are slightly different chemical structures at the same loci on homologous chromosomes. That is, they are simply chemically alternative versions of the same gene. Some genes have only one allele, while others have 20 or more.

FIGURE 3.18 Marfan Syndrome (a) The hand on the right shows normal finger growth. The hand on the left has much longer and thinner fingers due to Marfan syndrome, a hereditary disorder of the regulatory genes that control connective tissue. As a result of Marfan syndrome, uncontrolled bone growth leads to long and thin fingers and toes, long and thin arms and legs, and increased stature. Organs such as the lungs and heart can also be negatively affected. (b) In the 1960s, a scientific paper asserted that US president Abraham Lincoln (1809–1865) was afflicted with Marfan syndrome. This still- controversial assessment was based entirely on Lincoln’s unusual tallness and the length of his limbs.

(a)

(b)

 

 

62 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

Human blood type is one genetic trait with different alleles. Each person has one of four blood types— A, B, AB, or O— and these four types comprise the ABO blood group system, first discovered in 1900 (Figure 3.20). Because it has two or more vari- ants, a genetic trait such as this one is called a polymorphism (Greek poly, meaning “many”; Greek morph, meaning “form”). Each person has one A, B, or O allele on one chromosome of the homologous pair and another A, B, or O allele on the other chromosome of that pair. The combination determines the person’s blood type.

Although Mendel did not know about chromosomes, he recognized that phys- ical units of inheritance— which we now know to be the genes— segregate in a very patterned fashion. That is, his experiments with garden peas showed that the father contributes one physical unit and the mother contributes the other. This is Mendel’s law of segregation (Figure  3.21). For example, a person with blood type AB will pass on either an A or a B allele to a child but not both. The other allele will come from the other parent. This discovery was revolutionary because it explained how new variation arises in reproduction.

Some of the most exciting contemporary DNA research has revealed a whole new array of genetic markers. Showing a tremendous amount of variation within and between human populations, these SNPs are known from well over 1 million sites on the human genome. Closer examination of the human genome has also revealed that DNA segments are often repeated, sometimes many times and for no apparent reason. These repeated sections, or microsatellites, are highly

Chromosome

Antp gene

Fruit fly Hox genes Mouse Hox genes Goose Python Human

Hoxc6 gene

Neck

Thorax

Hox genes have been identified in all animals, plants, and fungi. They are found as a unique cluster known as the Hox cluster or Hox complex.

Unlike vertebrate animals, insects such as fruit flies have distinct body regions, including the head and the middle, or thorax. In fruit flies, the Hox gene that determines the thoracic region of the body during the larval stage of development is called Antp.

Other vertebrates, such as birds and reptiles, have a Hoxc6 gene, which determines the location of the thorax in the embryo. However, the location of the Hoxc6 gene varies with each animal, allowing for variation in the length of the cervical, or neck, region. If the Hoxc6 gene is lower in the body, as it is in geese, the animal will have a much longer neck than if the gene is located close to the head, as it is in pythons. Pythons, as a result of this placement, have virtually no neck.

While the body regions of vertebrates, such as mice, are not as distinct as those of flies and other insects, Hox genes determine their body regions during embryological development. The Hoxc6 gene in mice delimits the thoracic region, which is indicated by the thoracic vertebrae.

Humans, being vertebrates, also have a Hoxc6 gene, which determines the location of the thoracic region. Humans have a neck of intermediate length when compared to geese and pythons; the Hoxc6 gene is closer to the head in humans than in geese, but is lower than in pythons. The Hoxc6 gene is responsible not just for determining the location of the thorax; in humans, this gene determines the development of the entire thoracic region, including mammary glands.

FIGURE 3.19 Homeotic (Hox) Genes Discovered in 1983 by Swiss and American researchers, these regulatory genes are coded to produce proteins that turn on many other genes, in particular those that determine the regions of the body during prenatal development. Without these genes, or if there are mutations in these genes, body development may be altered. For example, a mutation in the Hox genes of a fruit fly can cause a leg instead of an antenna to grow from the head.

polymorphism Refers to the presence of two or more alleles at a locus and where the frequency of the alleles is greater than 1% in the population.

law of segregation Mendel’s first law, which asserts that the two alleles for any given gene (or trait) are inherited, one from each parent; during gamete produc- tion, only one of the two alleles will be present in each ovum or sperm.

microsatellites Also called short tandem repeats (STRs); refers to sequences of repeated base pairs of DNA, usually no more than two to six. If repeated exces- sively, they are often associated with neu- rological disorders, such as Huntington’s chorea.

 

 

individualistic, forming a unique DNA signature for each person. Microsatellites have quickly become the most important tool for individual identification, and they have proven especially valuable in forensic science. For example, they have been used to identify victims of the 9/11 attacks as well as genocide and mass- murder fatalities in the Balkans, Iraq, and Argentina.

GENOTYPES AND PHENOTYPES: GENES AND THEIR EXPRESSION The two alleles, whether they are chemically identical (e.g., AA) or chemically different (e.g., AO), identify the genotype— the actual genetic material in the pair of homologous chromosomes. Chemically identical alleles are called homozygous.

Red blood cells

Antigens

Antibody

Agglutination

(a)

(b)

FIGURE 3.20 Antibody–Antigen System When a person receives a blood transfusion, the transfused blood must have the same blood type as the recipient’s own to avoid an antibody– antigen reaction. (a) Each red blood cell has structures on its surface, known as antigens, that identify the cell as being type A, B, AB, or O. If the wrong type of blood is given to a person, the body’s immune system recognizes the new antigens as foreign. Special proteins called antibodies are then produced in the blood in response to the “invaders.” (b) The antibodies attach themselves to the foreign antigens, causing agglutination, or clumping, of the blood cells. Because the coagulated blood cannot pass through blood vessels properly, the recipient’s tissues do not receive the blood they need. If not treated, the person might die.

For all the blood types in the ABO blood group system, Table 3.2 shows the antigens, antibodies, and acceptable and unacceptable blood types. For example, type A blood, which can result from AO alleles or AA alleles, has A antigens on its surface and anti- B antibodies, which will react with B or AB blood. (Genotypes and phenotypes are defined and discussed below.)

antigens Specific proteins, on the surface of cells, that stimulate the immune sys- tem’s antibody production.

antibodies Molecules that form as part of the primary immune response to the presence of foreign substances; they attach to the foreign antigens.

homozygous Refers to the condition in which a pair of alleles at a single locus on homologous chromosomes are the same.

TABLE 3.2 The ABO Blood Group System

Phenotypes Genotypes Antigens Antibodies Unacceptable Blood Types

Acceptable Blood Types

A AO, AA A anti-B B, AB A, O

B BO, BB B anti-A A, AB B, O

AB AB A, B none none (universal recipient) A, B, AB, O

O OO none (universal donor) anti-A, anti-B A, B, AB O

Polymorphisms: Variations in Specific Genes | 63

 

 

64 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

Chemically different alleles are called heterozygous. When alleles are hetero- zygous, the dominant one will be expressed in the phenotype— the visible mani- festation of the gene. For example, individuals who are AA or who are AO have the same phenotype— A expresses dominance over O, and both individuals are blood type A. The recessive allele is not expressed. When you know simply that a person’s blood type is A, you cannot tell whether that person’s genotype is AA or  AO.  Rather, the blood type refers to the phenotype and not the genotype. If the person is AO, then the O allele is hidden because that allele is recessive. For the recessive allele to be expressed, each of the homologous chromosomes must have the recessive allele. For example, the alleles for type O blood are OO. For a

Pure red sweet peas (RR)

First (F1) Generation

Pure white sweet peas (rr)

R

100% Rr = red

r

r

Hybrid red (Rr)

Hybrid red (Rr)

R

Hybrid red (Rr)

Hybrid red (Rr)

Hybrid red (Rr)

Second (F2) Generation

Hybrid red (Rr)

R

25% RR = red, 50% Rr = red, 25% rr = white

R

r

Hybrid red (Rr)

Pure white (rr)

r

Pure red (RR)

Hybrid red (Rr)

The pure red parent has two R (red) alleles, so it can contribute only the R allele.

Each parent is a hybrid, with one R allele and one r allele.

The pure white parent has two r (white) alleles, so it can contribute only the r allele.

All the offspring will have Rr alleles. Since R (red) is dominant to r (white), all offspring will be red.

There is a 25% chance for an offspring plant with RR alleles (red), a 25% chance for one with rr alleles (white), and a 50% chance for one with Rr alleles (red).

FIGURE 3.21 Law of Segregation Mendel’s first law, the law of segregation, declares that the mother and father contribute equally to an offspring’s genetic makeup. For each gene, a person has two alleles (which can be the same or different). One allele is from the person’s mother, and one is from the person’s father.

Remember that meiosis (see Figure 3.13) creates four gametes, each of which has only one set of chromosomes, no pairs. Each gamete, having this one set, can pass on only one allele for each gene. If the gamete that the father contributes to fertilization has the allele for brown hair, for example, that is the only allele the father will contribute to the offspring. The other allele, for brown or a different color, will come from the mother.

heterozygous Refers to the condition in which a pair of alleles at a single locus on homologous chromosomes are different.

 

 

The Complexity of Genetics | 65

person to have type O blood, both alleles in the pair of homologous chromosomes have to be O.

Sometimes alleles exhibit codominance, where neither chemically different version dominates the other. In the ABO blood group system, the A and B alleles are codominant and both are expressed. If someone has type AB blood, you know that person’s phenotype and genotype.

The Complexity of Genetics Since the discovery of how the genetic code works (see chapter  2), the general impression about genes has been that they represent specific locations of DNA coded to produce specific proteins. Much of the field of genetics, including anthropological genetics, is based on this “one gene– one protein” model. How- ever, the relationship between genes and their physical expression turns out to be considerably more complex than previously thought. For example, many traits are polygenic, affected by genes at two or more loci. On the other hand, pleiotropy may be operating, whereby a single gene can have multiple effects. One example of pleiotropic effects is Marfan syndrome (see Figure  3.18). This condition is a dominant collagen disorder caused by a gene on chromosome 15. Collagen is a major protein, representing 25%–35% of all protein in the body. Because colla- gen is found in so many types of body tissues— for example, skeletal, visual, and cardiovascular— this single gene’s effects are pervasive.

The physical manifestations also may be influenced by environmental factors. In humans, thousands of complex phenotypes— such as birth weight, height, skin color, head form, tooth size, and eye shape— have multiple genetic components and are influenced by environmental factors. Many other complex phenotypes— such as autism spectrum and other behavioral and cognitive disorders— may have envi- ronmental influences.

Scientists have long understood that environment plays an important role in the physical manifestations of various aspects of one’s genome. For example, environmental factors affecting the mother can also affect the developing fetus. If the fetus is female, then those offspring resulting from the developing ova of that fetus may also be affected. In this way, environmental factors operating at a given point in time can affect the health and well- being of subsequent generations. Such epigenetic effects represent potentially heritable changes in behavior or biology but without alteration in the DNA sequence. These changes alter the way that DNA is regulated, without altering the DNA itself. A new and exciting area of research is showing that epigenetic mechanisms occurring within cells may be activated by a variety of behavioral and environmental factors, such as smoking, alcohol consumption, nutrition, physical activity, temperature extremes, and disease. In this regard, regulation of DNA in the offspring may be altered due to epigenetic phenomena contributing to birth defects and other outcomes.

The interaction between the environment and the genes of the mother and her embryo, from conception to birth, has considerable impact on the normal growth of the fetus as well as risk for disease following birth. Indeed, obesity and some birth defects, common diseases, and cancers appear to be influenced by epigene- tic factors. For example, key nutrients— including vitamins A, B3, C, and D— play roles in regulating DNA. An insufficiency of these nutrients is linked to diabetes, atherosclerosis, and cancer. By investigating the epigenetics of health and behavior, scientists have opened up a pathway toward understanding factors that influence

codominance Refers to two different alleles that are equally dominant; both are fully expressed in a heterozygote’s phenotype.

polygenic Refers to one phenotypic trait that is affected by two or more genes.

epigenetic Refers to heritable changes but without alteration in the genome.

 

 

66 | CHAPTER 3 Genetics: Reproducing Life and Producing Variation

parent and offspring health. In short, researchers have helped show how environ- ment interacts with the genome across generations.

For some phenotypic traits, scientists can determine only the relative pro- portions of genetic and environmental contributions. A trait’s heritability, the proportion of its variation that is genetic, can be calculated this way:

genetic variation heritability (H2) = (genetic variation + environmental variation)

Heritability estimates are presented as values ranging from 0, where none of the variation is genetic in origin, to 1, where all of the variation is genetic. In traits with heritability estimates greater than .5, most of the variation is genetic. For example, in the United States, the heritabilities for height and weight are estimated to be .6 and .3, respectively. Tooth size has among the highest levels of heritability, about .7, and brain size and fingerprints are even higher, at .9. Physical anthropologists and other evolutionary biologists are very interested in heritability for one simple reason: because only heritable traits respond to natural selection, they are the primary driving force of evolution.

Measurement of heritability, however, is complicated by pleiotropy— a single allele can have multiple effects. In fact, most complex traits are polygenic and pleiotropic (Figure 3.22).

The DNA Revolution has made it possible to understand in much greater detail the underlying principles of inheritance laid out so eloquently by Gregor Mendel a century and a half ago. In presenting the great breadth of knowledge about how genetic variation is transmitted from parents to offspring and maintained devel- opmentally within individuals, this chapter has laid the groundwork for the topic of the next chapter, the study of genetic change in populations.

heritability The proportion of phenotypic variation in a population that is due to genetic variation across individuals rather than variation in the environmental condi- tions experienced by the individuals. This proportion can vary from one population to another, and thus it provides a sense of the contribution of genetic influence for each population.

Each gene has a distinct biological effect.

Gene Effect

Polygenic trait: many genes contribute to a single effect.

Gene Effect

Pleiotropy: one gene has multiple biological effects.

(a) (b)

(c) (d)

Gene Effect

Polygenic traits and pleiotropy.

Gene Effect

FIGURE 3.22 Polygenic Traits and Pleiotropic Genes (a) Mendel’s simple rules of inheritance (b) do not apply when one trait is affected by many genes. Eye color, for example, is determined by at least three genes. Because this trait is polygenic, some children’s eye colors are very different from their parents’. (c) Pleiotropic genes affect more than one physical trait. The PKU allele, for example, affects mental abilities and the coloration of hair and skin. A person who inherits this allele will have the disease phenylketonuria, in which a missing enzyme leads to mental retardation as well as reduced hair and skin pigmentation. (d) One trait can be affected by several genes, and each of those genes can affect several other traits as well.

 

 

67

A N S W E R I N G T H E B I G Q U E S T I O N S

C H A P T E R   3 R E V I E W

What is the genetic code? • The genetic code is DNA, packaged in individual

chromosomes. • One type of DNA exists in the cell’s nucleus (nuclear

DNA), and the other type exists in the cytoplasm (mitochondrial DNA). Nuclear DNA provides most of the genetic code.

What does the genetic code (DNA) do? • DNA serves as the chemical template for its own

replication. DNA replication is the first step toward the production of new cells: somatic cells and gametes. — Mitosis results in the production of two identical

somatic cells. In humans, each parental or daughter cell has 46 chromosomes in 23 homologous pairs.

— Meiosis results in the production of four gametes. In humans, each parental cell has 46 chromosomes, and each gamete has 23 chromosomes.

• DNA serves as the chemical template for the creation of proteins. — Proteins are combinations of amino acids. They

are responsible for all physical characteristics (structural proteins), the regulation (regulatory or functional proteins) of activities within cells (enzymes), the regulation of activities between cells (hormones), and the fighting of foreign antigens (antibodies). Thus, proteins comprise the entire body and determine all of its functions, from conception through maturity.

— The two types of proteins are governed by the two corresponding types of genes, structural and regulatory.

— Hox genes are regulatory genes that control the development of body parts, such as limbs and internal organs, and their locations relative to each other.

What is the genetic basis for human variation? • Each chromosome has a linear sequence of

nucleotides that are coded to produce specific bodily structures and functions. These linear sequences are genes, and each gene has a particular locus on each chromosome— and the same locus on the like (homologous) chromosome.

• Each pair of homologous chromosomes consists of a paternal chromosome and a maternal chromosome.

• An individual’s genotype, or actual genetic composition, is identified on the basis of two alleles, one from the father and one from the mother. Alleles can be chemically identical or chemically different. The genotype is expressed physically as a phenotype.

• For most traits, no direct map yet exists for the translation of genotype to phenotype. That is, one gene may result in the construction of one functional protein, or multiple polypeptides each produced by a different gene may form the functional protein. The genetic basis for such a protein, therefore, is difficult to determine.

• Most physical characteristics are determined by more than one gene (polygenic), and some genes can have multiple effects (pleiotropy).

• Epigenetics is the source of a new revolution in our understanding of how genes work. By viewing the profound impact of environment on gene function, an impact created by both biological and cultural circumstances, we can gain new insight into inheritance, health, and behavior.

1

2

3

REVIEW THIS CHAPTER WITH PERSONALIZED, INTERACTIVE QUESTIONS THROUGH IJK, INQUIZITIVE.WWNORTON.COM q

 

 

68

K E Y T E R M S adenine adenosine triphosphate (ATP) amino acids antibodies anticodons antigens autosomes coding DNA codominance codons complementary bases crossing- over cytoplasm cytosine diploid cell epigenetic essential amino acids eukaryotes free- floating nucleotides gametes genome guanine haplogroups haploid cell haplotypes heritability

heteroplasmic heterozygous homeotic (Hox) genes homologous homoplasmic homozygous karyotype law of independent assortment law of segregation linkage locus matriline meiosis messenger RNA (mRNA) microsatellites mitochondria mitosis monosomy noncoding DNA nondisjunctions nucleotide nucleus paleogenetics patriline peptide bond polygenic

polymerase chain reaction (PCR) polymorphism polypeptide prokaryotes recombination regulatory genes regulatory proteins replication ribonucleic acid (RNA) ribosomal RNA (rRNA) ribosomes sex chromosomes single nucleotide polymorphisms (SNPs) somatic cells structural genes structural proteins thymine transcription transfer RNA (tRNA) translation translocations triplets trisomy uracil zygote

E V O L U T I O N R E V I E W Insights from Genetics

Synopsis DNA is often described as a genetic blueprint as it encodes the plan for an organism’s traits and ensures that these traits are passed on to future generations. One of the major func- tions of the DNA molecule is replication: creating identical copies of itself. The reliability of this process influences evolution by making variation heritable—that is, traits will be passed on to offspring, who will be similar to their parents. The second major function of DNA is directing protein synthesis, which is how genotype (genetic code) is translated to phenotype (physical expression of this code). Protein synthesis is the basis for the traits that allow organisms to interact with their environment and undergo natural selection. As our knowledge of molecular genetics has expanded, so has our appre- ciation for its complexity. Many traits are polygenic (influenced by

more than one gene), and many genes are pleiotropic (operating on more than one trait). Furthermore, epigenetic (environmental, nonge- netic) phenomena can alter DNA regulation and phenotype without altering the DNA sequence itself. Thus, genetics both complicates and illuminates our understanding of human variation and evolution.

Q1. What are the two types of cell division? How does each type affect individual genetic variation?

Q2. What is recombination? How does the effect of recombination differ from that of mutation or a random change in the DNA sequence?

Q3. Given the modern “DNA Revolution” and our growing know- ledge of humans from a genetic perspective, why is it

 

 

69

still important for physical anthropologists to study physical remains (e.g., bones and teeth) when testing hypotheses about evolution and variation among ourselves and our closest living and fossil relatives?

Hint Think about the types of remains available for living and fossil organisms and about how genetics are related to physical characteristics.

Q4 . Consider various human characteristics, such as height, weight, skin color, head form, and eye shape. Are the phe- notypic expressions of these characteristics discontinuous

(able to be assigned to discrete categories) or continuous (on a continuum rather than separable into discrete categories)? What might they tell us about the mechanisms behind human variation?

Q5. Epigenetic phenomena are environmental factors that affect the way DNA is regulated and may affect future generations. How do epigenetic phenomena differ from Lamarckian inheritance of acquired characteristics (discussed in the previous chapter)?

Hint Compare the heritability of characteristics involved in epigenetics and Lamarckism.

A D D I T I O N A L R E A D I N G S

Carey, N. 2012. The Epigenetics Revolution: How Modern Biology Is Rewriting Our Understanding of Genetics, Disease, and Inheritance. New York: Columbia University Press.

Kaestle, F. A. 2010. Paleogenetics: Ancient DNA in anthropology. Pp. 427–441 in C. S. Larsen, ed. A Companion to Biological Anthro- pology. Chichester, UK: Wiley- Blackwell.

Mielke, J. H., L. W. Konigsberg, and J. H. Relethford. 2006. Human Biological Variation. New York: Oxford University Press.

Portugal, F. H. and J. S. Cohen. 1977. A Century of DNA: A History of the Discovery of the Structure and Function of the Genetic Sub- stance. Cambridge, MA: MIT Press.

Relethford, J. H. 2003. Reflections of Our Past: How Human History Is Revealed in Our Genes. Boulder, CO: Westview Press.

Sapolsky,  R.  M.  2004. Of mice, men, and genes. Natural History May: 21–24, 31.

Sykes,  B.  2001. The Seven Daughters of Eve: The Science That Reveals Our Genetic Ancestry. New York: Norton.

Weiss,  M.  L.  and  J.  Tackney. 2012. An introduction to genetics. Pp. 53–98 in S. Stinson, B. Bogin, and D. O’Rourke, eds. Human Biology: An Evolutionary and Biocultural Perspective, 2nd ed. Hobo- ken, NJ: Wiley- Blackwell.

E V O L U T I O N R E V I E W

 

 

THE KEY DRIVER OF EVOLUTION is natural selection. Members of species having genetic variation that enhances survival to reproductive age will tend to live to reproductive age. In these three very different animals— the leafy sea dragon, deer mouse, and peppered moth— there is selection for genes con- trolling for pigmentation and other attributes of body phenotype. These pheno- types make it difficult for predators to see these prey. This “visual” selection is commonplace worldwide now, as it must have been in the past.

 

 

71 71

4 What causes evolutionary (genetic) change?

How is evolutionary (genetic) change measured, and how is the cause determined?

Genes and Their Evolution Population Genetics

One of the great stories about genetics comes not from a research program of a famous scientist in charge of a large laboratory filled with technicians, grad-uate students, and postdoctoral fellows and funded by a multimillion- dollar grant but rather from a student with a simple hypothesis and a passionate interest in testing a hypothesis. At Oxford University in the late 1940s, a 20-something Anthony Allison was finishing his coursework in basic sciences and was about to start his clinical medical training. Allison had grown up in Kenya; his interest in anthropology reflected his intellectual curiosity, but his desire to become a doctor was motivated by an ambi- tion to help improve native Kenyans’ quality of life. While at Oxford, he was exposed to the ideas of the English scientists R. A. Fisher and J. B. S. Haldane and the American scientist Sewall Wright, pioneers of the new field of population genetics and advocates of the novel idea that gene frequencies were tied to natural selection.

Following a bout of malaria, Allison decided to help Kenyans (and other peoples) by seeking a cure for this disease. In 1949, he joined an expedition to document blood groups and genetic traits in native Kenyans. On this expedition, Allison discov- ered that in areas affected by malaria, especially along Kenya’s coast (southeast) and near Lake Victoria (southwest), a remarkably high 20%–30% of the population carried the gene for sickle- cell anemia. But in the highlands (west), where there was no malaria, less than 1% of the people carried the gene. In what he described as a “flash of inspiration,” he hypothesized that individuals with the sickle- cell allele were resistant to malaria and that natural selection was operating on the gene. But how? he wondered.

1

2

B I G Q U E S T I O N S ?

 

 

72 | CHAPTER 4 Genes and Their Evolution: Population Genetics

Before pursuing these ideas, Allison completed his medical education. In 1953, with just enough money to buy passage back to Kenya and for food and simple lodging, he spent a year researching the relationship between malaria and sickle- cell anemia. He generated lab and field data, looking at malaria infection rates in people with and without the sickle- cell allele. His results showed that carriers of the gene are much more likely to survive malaria than are noncarriers. Natural selection was favoring the carriers.

The next year, Allison published three landmark scientific articles in rapid succes- sion, laying out the proof for his hypothesis. Gene frequencies are tied to natural selec- tion: carriers of the sickle- cell allele survive longer and produce more offspring than do noncarriers. In Allison’s words, “disease is an agent of natural selection.” Although his ideas met with strong skepticism, eventually Allison’s hypothesis was accepted by most scientists. His research enabled generations of geneticists and anthropologists to further investigate genes and their evolution, all of these researchers asking why some gene frequencies remain the same while others change over time.

Human populations exhibit some remarkable biological differences, and the sci- ence of genetics helps biologists answer questions about those differences. The questions and the answers are founded on Darwin’s discovery that phenotypes— the physical manifestations of genes— change over time. In addition, Mendel’s research on garden peas revealed how the inheritance of genes produces variation in phe- notypes. These two revolutionary scientific discoveries inform our understanding of biological variation and its evolution.

Before considering this chapter’s big question, we need to look at populations and species, the units that evolutionary biologists work with.

Demes, Reproductive Isolation, and Species To show how genetic variation is produced, the previous chapter focused on the individual and the transmission of genes from parents to offspring. When phys- ical anthropologists and geneticists study the genetics of individuals, they focus on the reproductive population, or deme: members of a species that produce offspring. That is, evolution is about groups of organisms that have the potential to reproduce. When physical anthropologists talk about populations, they often refer to the gene pool, which is all the genetic material within a population. When geneticists talk about the gene pool, they are even more specific, referring to all the variation within a specific genetic locus. For example, some people carry the sickle- cell allele and some do not.

The concept of the breeding population is also central to the definition of species. A species is comprised of all the populations (and their individual mem- bers) that are capable of breeding with each other and producing viable (fertile) offspring. Species, therefore, are defined on the basis of reproductive isolation (Figure  4.1). In biological terms, if two populations are reproductively isolated, members of one population cannot interbreed with members of the other. Repro- ductive isolation is largely related to geographic isolation. If two populations of the same species become isolated, such as by a mountain range or a large body

deme A local population of organisms that have similar genes, interbreed, and produce offspring.

gene pool All the genetic information in the breeding population.

reproductive isolation Any circumstance that prevents two populations from interbreeding and exchanging genetic material, such as when two populations are separated by a large body of water or a major mountain range.

 

 

Demes, Reproductive Isolation, and Species | 73

of water, enough  genetic differences could accumulate for two entirely different species to emerge.

In the living world, we can observe members of a species to verify that they can produce offspring. Obviously, we cannot do this with fossils. Rather, we have to infer reproductive isolation in fossil populations on the basis of geographic dis- tance and the physical resemblance between fossils. Fossil remains that share the same characteristics in morphology of teeth and bones likely represent members of the same species (Figure 4.2). (This important concept will inform the discussion of primate evolution in chapters 9–13. Fossils are the subject of chapter 8.)

(a)

(b) (c)

FIGURE 4.1 Reproductive Isolation (a) This map depicts the distribution of two related bird species: (b) ostriches are found in Africa, while (c) emus are found in Australia. These two species share a common ancestor, but the ocean between the two groups prevented them from interbreeding for thousands of years. Eventually, this geographic isolation led to reproductive isolation and two separate species.

 

 

74 | CHAPTER 4 Genes and Their Evolution: Population Genetics

Population genetics (see “The Evolutionary Synthesis, the Study of Popula- tions, and the Causes of Evolution” in chapter 2) is the study of changes in genetic material— specifically, the change in frequency of alleles (genes). Genes are the records from which evolution is reconstructed, both over the course of a few genera- tions (microevolution) and over many generations (macroevolution; Figure 4.3). Geneticists strive to document genetic change and to explain why it occurred. Such documentation and explanation are the central issues of evolutionary biology. Pop- ulation geneticists, physical anthropologists, and other evolutionary biologists tend to focus on genetic change over time. For example, if a trait in a population has two alleles, A and a, and the parent generation is 60% A and 40% a and the next generation is 65% A and 35% a, scientists would want to know why the evolution occurred— why the frequency of the A allele increased in the population.

Just as interesting, however, are instances in which frequency does not change over time— that is, when the frequencies of a population’s alleles for a particular

FIGURE 4.2 Living Fossils Fossilized remains might represent animals and plants that lived thousands or millions of years ago, but a number of such organisms have counterparts that live today. Compare the forms of the following organisms: (a) living horseshoe crab and (b) fossil of Mesolimulus walchii, the ancestor to the modern horseshoe crab; (c) living cockroach and (d) a 49  million– year- old fossil cockroach; (e) living American opossum (“playing possum,” i.e., feigning death) and (f) fossil skull of Didelphis albirentris, an ancestor to modern opossums. Which features have been maintained and which have been lost?

(a) (c)

living fossil

(b) (d)

(e) (f)

microevolution Small- scale evolution, such as changes in allele frequency, that occurs from one generation to the next.

macroevolution Large- scale evolution, such as a speciation event, that occurs after hundreds or thousands of generations.

FIGURE 4.3 Microevolution and Macroevolution (a) Microevolution is change in gene frequency over a few generations. For example, the beetles in this diagram have either of two colors, represented by two alleles. In the first generation, 75% of the color alleles are green and 25% are brown. In the second generation, 71% of the color alleles are green and 29% are brown. Thus, the frequencies of the green and brown alleles have undergone a microevolutionary change. (b) Macroevolution is substantial change over many generations, the creation of new species. For example, over 60 million years, the eohippus— a small, dog- sized animal, with multitoed feet, that inhabited rainforests— evolved into the modern horse. Among the species’ large- scale changes were increases in overall body size and height as well as the loss of toes. Horses’ single- toed hooves enable them to run more efficiently in the open grasslands they naturally inhabit.

First generation

75% 25%

Second generation

×

×

×

Modern horse

Pliohippus

Merychippus

Mesohippus

Eohippus

height 1.6 m (5.25 ft)

height 1.0 m (3.28 ft)

height 1.0 m (3.28 ft)

height .6 m (1.97 ft)

height .4 m (1.31 ft)

1 m

ya 10

m ya

30 m

ya 40

m ya

60 m

ya

71% 29%

(b)

(a)

 

 

Demes, Reproductive Isolation, and Species | 75

Population genetics (see “The Evolutionary Synthesis, the Study of Popula- tions, and the Causes of Evolution” in chapter 2) is the study of changes in genetic material— specifically, the change in frequency of alleles (genes). Genes are the records from which evolution is reconstructed, both over the course of a few genera- tions (microevolution) and over many generations (macroevolution; Figure 4.3). Geneticists strive to document genetic change and to explain why it occurred. Such documentation and explanation are the central issues of evolutionary biology. Pop- ulation geneticists, physical anthropologists, and other evolutionary biologists tend to focus on genetic change over time. For example, if a trait in a population has two alleles, A and a, and the parent generation is 60% A and 40% a and the next generation is 65% A and 35% a, scientists would want to know why the evolution occurred— why the frequency of the A allele increased in the population.

Just as interesting, however, are instances in which frequency does not change over time— that is, when the frequencies of a population’s alleles for a particular

FIGURE 4.2 Living Fossils Fossilized remains might represent animals and plants that lived thousands or millions of years ago, but a number of such organisms have counterparts that live today. Compare the forms of the following organisms: (a) living horseshoe crab and (b) fossil of Mesolimulus walchii, the ancestor to the modern horseshoe crab; (c) living cockroach and (d) a 49  million– year- old fossil cockroach; (e) living American opossum (“playing possum,” i.e., feigning death) and (f) fossil skull of Didelphis albirentris, an ancestor to modern opossums. Which features have been maintained and which have been lost?

(a) (c)

living fossil

(b) (d)

(e) (f)

microevolution Small- scale evolution, such as changes in allele frequency, that occurs from one generation to the next.

macroevolution Large- scale evolution, such as a speciation event, that occurs after hundreds or thousands of generations.

FIGURE 4.3 Microevolution and Macroevolution (a) Microevolution is change in gene frequency over a few generations. For example, the beetles in this diagram have either of two colors, represented by two alleles. In the first generation, 75% of the color alleles are green and 25% are brown. In the second generation, 71% of the color alleles are green and 29% are brown. Thus, the frequencies of the green and brown alleles have undergone a microevolutionary change. (b) Macroevolution is substantial change over many generations, the creation of new species. For example, over 60 million years, the eohippus— a small, dog- sized animal, with multitoed feet, that inhabited rainforests— evolved into the modern horse. Among the species’ large- scale changes were increases in overall body size and height as well as the loss of toes. Horses’ single- toed hooves enable them to run more efficiently in the open grasslands they naturally inhabit.

First generation

75% 25%

Second generation

×

×

×

Modern horse

Pliohippus

Merychippus

Mesohippus

Eohippus

height 1.6 m (5.25 ft)

height 1.0 m (3.28 ft)

height 1.0 m (3.28 ft)

height .6 m (1.97 ft)

height .4 m (1.31 ft)

1 m

ya 10

m ya

30 m

ya 40

m ya

60 m

ya

71% 29%

(b)

(a)

 

 

76 | CHAPTER 4 Genes and Their Evolution: Population Genetics

trait are in a state of equilibrium. For example, in areas of West Africa where malaria is common, the frequency of the sickle- cell allele remains relatively con- stant. What factors— forces of evolution— account for deviations from equilibrium?

Hardy- Weinberg Law: Testing the Conditions of Genetic Equilibrium In 1908, Godfrey Hardy (1877–1947), an English mathematician, and Wilhelm Weinberg (1862–1937), a German obstetrician, independently recognized that some alleles are in a state of equilibrium. If no mutation or natural selection or gene flow occurs, if the population is large, if mating is random, and if all members of the population produce the same number of offspring, then genotype frequencies at a single gene locus will remain the same after one generation. Moreover, the equilibrium frequencies will be a function of the allele frequencies at the locus. This is called the Hardy- Weinberg law of equilibrium. In the simplest case (Table 4.1), a single locus has A (dominant) and a (recessive) alleles, with respective frequencies of p and q. In assessing the population as a whole, it is assumed that males and females have both alleles. The Hardy- Weinberg law predicts the geno- type frequencies for the next generation after one mating, where p2 is the genotype frequency for the AA homozygous alleles, 2pq is the genotype frequency for the Aa (heterozygous) alleles, and q2 is the genotype frequency for the aa homozygous alleles. In other words, the total population (100%) should be the sum of the frequencies of three genotypes, expressed by the simple mathematical equation p2 + 2pq + q2 = 1. If a hypothetical population were 60% A ( p = .6) and 40% a (q = .4), then the genotype frequencies in the next generation would work out to AA = .36, Aa = .48, and aa = .16. The frequencies can be expressed as decimals or percentages, but they are expressed most often as decimals. Since the three genotypes are the only genotypes for the gene in question in the population, the frequencies must add up to 1 or 100%. So, if the frequency of AA is .36 (or 36%), the frequency of Aa is .48 (or 48%), and the frequency of aa is .16 (or 16%), together they add up to 1 (or 100%).

In the absence of evolution, the frequencies of the genotypes will in theory remain the same forever. In this way, the Hardy- Weinberg equilibrium hypothe- sizes that gene frequencies remain the same because no evolutionary change takes place (Figure 4.4).

By determining the genotype frequencies for a population at different points in time, however, the Hardy- Weinberg equation establishes expectations as to whether evolution is operating on a particular gene. If the genotype frequencies

equilibrium A condition in which the sys- tem is stable, balanced, and unchanging.

TABLE 4.1 Punnett Square for Hardy- Weinberg Equilibrium

Females

A (p) a (q)

Males A (p) AA (p2) Aa (pq)

a (q) Aa (pq) aa (q2)

Hardy- Weinberg law of equilibrium A math- ematical model in population genetics that reflects the relationship between frequencies of alleles and of genotypes; it can be used to determine whether a population is undergoing evolutionary changes.

 

 

Mutation: The Only Source of New Alleles | 77

change from one generation to the next, the population is not in equilibrium— it is evolving. If the frequencies remain the same, the population is in equilibrium— the population is not evolving, at least with respect to the locus being studied.

What might cause a population to change its allele frequencies and go out of equilibrium? As noted in chapters  2 and 3, genes are passed from generation to generation by interbreeding within populations in particular and among members of the same species in general, and genetic changes result from one or a combina- tion of the four forces of evolution: mutation, natural selection, genetic drift, and gene flow.

Mutation: The Only Source of New Alleles During cell reproduction, DNA almost always replicates itself exactly. Sometimes, however, the replication process produces an error or a collection of errors in the DNA code. If the problem is not at once detected and corrected by a set of enzymes that monitor DNA, a mutation results. The mutation can be any heritable change in the structure or amount of genetic material.

FIGURE 4.4 Gene Frequencies in Equilibrium As this Punnett square illustrates, the Hardy- Weinberg equilibrium captures gene frequencies in a static moment, when no evolutionary change is taking place. Crossing the males (sperm pool) and females (egg pool) of the population produces theoretical genotype frequencies of the next generation.

Once evolutionary change takes place, the actual genotype frequencies will differ significantly from the theoretical genotype frequencies expressed in this Punnett square. For example, if the population later turns out to be 5% AA, 65% aa, and 30% Aa, an evolutionary force has most likely altered its genotype frequencies.

A p = .6

Sperm pool

a q = .4

A p = .6

a q = .4

Aa pq .24

Aa pq .24

Eg g

po ol

AA p2

0.36

aa p2

0.16

AA p2

.36

aa q2

.16

The allele frequencies for males and females in this population are the same: p (frequency of A allele) is .6 and q (frequency of a allele) is .4.

Approximately 36% of the offspring will have the genotype AA. The estimate is made by multiplying the male p (frequency of A allele = .6) with the female p (.6). Thus, p × p = p2, and .6 × .6 = .36. The frequency of the AA genotype is represented by p2.

Approximately 48% of the offspring population will have the genotype Aa. This estimate is made by multiplying the male p (.6) with the female q (.4) and the male q (.4) with the female p (.6). Thus, p × q = pq and q × p = qp; .6 × .4 = .24 and .4 × .6 = .24. Since both p × q and q × p must be included, the two results are added together. 2pq = 2 × .24 (or .24 + .24) = .48. The frequency of the Aa genotype is represented by 2pq.

Approximately 16% of the offspring will have the geno- type aa. This estimate is made by multiplying the male q (frequency of a allele = .4) with the female q (.4). Thus, q × q = q2, and .4 × .4 = .16. The frequency of the aa genotype is represented by q2.

 

 

78 | CHAPTER 4 Genes and Their Evolution: Population Genetics

Because so much of any person’s DNA is noncoding (see “Producing Proteins: The Other Function of DNA” in chapter 3), many mutations do not affect the indi- vidual’s health, well- being, or survival. A new sequence of coding DNA that results from mutation may have profound consequences, positive or negative. For example, the mutation might code the DNA for a protein with an altered or different func- tion from that performed by the protein coded for in the original parent strand of DNA, or the mutation might create a sequence that results in either no protein or an abnormal protein (Figure 4.5). Mutations occur at random and can occur in any cell, but the ones with consequences for future generations take place in gametes. Gametes may transfer mutations to offspring, depending on what happens during meiosis in the parents. Regardless of their causes or outcomes, mutations are the only source of new genetic variation in a population.

Mutations involving incorrect base pairing are called point mutations. A syn- onymous point mutation creates an altered triplet in the DNA, but the alteration carries with it the original amino acid. Because the amino acid is the same, the pro- tein formed is the same. A nonsynonymous point mutation results in a matchup that brings along a different amino acid. Such a mutation can have dramatic results for the individual carrying it. For example, a mutation on human chromosome 11 results in a GUG codon instead of a GAG codon. The GUG codon is encoded to produce the amino acid valine, whereas the GAG codon would have normally led to the production of glutamic acid. This substitution leads to the abnormal hemo- globin that results in sickle- cell anemia (discussed later in this chapter).

As a result of the shifting base pairs caused by base insertion, the reading frame of a gene is altered or stopped entirely. This frameshift mutation produces a protein having no function. Such a mutation usually involves a small part of the DNA sequence, often just a base pair or a relatively limited number of base pairs.

Other kinds of mutations can affect far more of the genome. Transposable ele- ments are genes that can copy themselves to entirely different places along the DNA sequence. If such a gene inserts itself into another gene, it can fundamentally alter the other gene, doing real damage. If, as is strongly likely, the gene transposes itself to a noncoding area of the DNA sequence, little or no significant alteration will occur.

Large parts of DNA sequences or entire chromosomes can be affected by muta- tions. An entire piece of chromosome can be moved to another chromosome, can be placed differently on the same chromosome, or can be positioned in a chromo- some backward. The impacts of these mutations are highly variable and depend on the mutations’ loci.

In the most extreme mutations, entire chromosomes can be duplicated (a tri- somy) or lost altogether (a monosomy). Examples of trisomies are Down syndrome, with its extra twenty- first chromosome (see “Meiosis: Production of Gametes [Sex Cells]” in chapter  3) and Klinefelter’s syndrome, a common sex chromosome variant that appears in about 1 of 500–1,000 births.

All mutations fall into either of two types: spontaneous mutations have no known cause (Figure 4.6); induced mutations are caused by specific environmen- tal agents, usually associated with human activity. These agents, or mutagens, are increasingly becoming known. For example, ionizing radiation ( X- rays) and various toxic chemicals have been linked to mutations in animals and humans. Most mutations are spontaneous, however, and are simply DNA copying errors. The human mutation rate is higher in male sex cells (sperm) than in female sex cells (eggs) but is generally on the order of one per million per nucleotide per generation. The human genome includes about 3 billion base pairs, about 1.5% of which are protein- coding, so the average mutation rate in humans is .45 mutations

point mutations Replacements of a single nitrogen base with another base, which may or may not affect the amino acid for which the triplet codes.

synonymous point mutation A neutral point mutation in which the substituted nitrogen base creates a triplet coded to produce the same amino acid as that of the original triplet.

nonsynonymous point mutation A point mutation that creates a triplet coded to produce a different amino acid from that of the original triplet.

frameshift mutation The change in a gene due to the insertion or deletion of one or more nitrogen bases, which causes the subsequent triplets to be rearranged and the codons to be read incorrectly during translation.

transposable elements Mobile pieces of DNA that can copy themselves into entirely new areas of the chromosomes.

Klinefelter’s syndrome A chromosomal trisomy in which males have an extra X chromosome, resulting in an XXY con- dition; affected individuals typically have reduced fertility.

spontaneous mutations Random changes in DNA that occur during cell division.

induced mutations Refers to those muta- tions in the DNA resulting from exposure to toxic chemicals or to radiation.

mutagens Substances, such as toxins, chemicals, or radiation, that may induce genetic mutations.

T

A

Base substitution

(a) Base substitution

(b) Base insertion

A base substitution, the simplest kind of mutation, occurs when a single nitrogen base is substituted by another base. Here, thymine is being replaced by cytosine.

T

A

U

A

U

A

U

A

U

T

A

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

C

G

C

C

G

C

G

C

G

C

Valine Serine Serine Valine Proline

Base insertion

A

U

A

U

A

U

A

U

A

A

U

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

CG

C

G

C

Valine Serine Serine Tyrosine Serine

Normal gene

Mutation

Mutation

DNA template

mRNA

Protein

A

U

A

U

A

U

A

U

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

C

G

CG

C

G

C

Valine Serine Serine Isoleucine Proline

The mRNA receives the appropriate complementary nitrogen base, uracil. When the mRNA reaches the ribosomes, the series of codons is changed because of the insertion. Instead of reading the correct AUU, the ribosomes read UAU, which codes for the amino acid tyrosine instead of isoleucine. Again, this amino acid change may or may not affect the resulting protein.

A base insertion occurs when a nitrogen base is added to the DNA template. Here, adenine is inserted.

As a result of the base substitution, the mRNA strand receives complementary guanine rather than the adenine that pairs with thymine. This further substitution changes the amino acid that is added to the polypeptide chain at the ribosomes. Instead of isoleucine, valine is inserted (see Table 3.2 in chapter 3). This amino acid change may or may not affect the resulting protein.

FIGURE 4.5 DNA Mutations During the transcription phase of protein synthesis, errors in the DNA template can affect the resulting protein. Two types of DNA mutation are illustrated in these diagrams: (a) base substitutions and (b) base insertions.

 

 

Mutation: The Only Source of New Alleles | 79

Because so much of any person’s DNA is noncoding (see “Producing Proteins: The Other Function of DNA” in chapter 3), many mutations do not affect the indi- vidual’s health, well- being, or survival. A new sequence of coding DNA that results from mutation may have profound consequences, positive or negative. For example, the mutation might code the DNA for a protein with an altered or different func- tion from that performed by the protein coded for in the original parent strand of DNA, or the mutation might create a sequence that results in either no protein or an abnormal protein (Figure 4.5). Mutations occur at random and can occur in any cell, but the ones with consequences for future generations take place in gametes. Gametes may transfer mutations to offspring, depending on what happens during meiosis in the parents. Regardless of their causes or outcomes, mutations are the only source of new genetic variation in a population.

Mutations involving incorrect base pairing are called point mutations. A syn- onymous point mutation creates an altered triplet in the DNA, but the alteration carries with it the original amino acid. Because the amino acid is the same, the pro- tein formed is the same. A nonsynonymous point mutation results in a matchup that brings along a different amino acid. Such a mutation can have dramatic results for the individual carrying it. For example, a mutation on human chromosome 11 results in a GUG codon instead of a GAG codon. The GUG codon is encoded to produce the amino acid valine, whereas the GAG codon would have normally led to the production of glutamic acid. This substitution leads to the abnormal hemo- globin that results in sickle- cell anemia (discussed later in this chapter).

As a result of the shifting base pairs caused by base insertion, the reading frame of a gene is altered or stopped entirely. This frameshift mutation produces a protein having no function. Such a mutation usually involves a small part of the DNA sequence, often just a base pair or a relatively limited number of base pairs.

Other kinds of mutations can affect far more of the genome. Transposable ele- ments are genes that can copy themselves to entirely different places along the DNA sequence. If such a gene inserts itself into another gene, it can fundamentally alter the other gene, doing real damage. If, as is strongly likely, the gene transposes itself to a noncoding area of the DNA sequence, little or no significant alteration will occur.

Large parts of DNA sequences or entire chromosomes can be affected by muta- tions. An entire piece of chromosome can be moved to another chromosome, can be placed differently on the same chromosome, or can be positioned in a chromo- some backward. The impacts of these mutations are highly variable and depend on the mutations’ loci.

In the most extreme mutations, entire chromosomes can be duplicated (a tri- somy) or lost altogether (a monosomy). Examples of trisomies are Down syndrome, with its extra twenty- first chromosome (see “Meiosis: Production of Gametes [Sex Cells]” in chapter  3) and Klinefelter’s syndrome, a common sex chromosome variant that appears in about 1 of 500–1,000 births.

All mutations fall into either of two types: spontaneous mutations have no known cause (Figure 4.6); induced mutations are caused by specific environmen- tal agents, usually associated with human activity. These agents, or mutagens, are increasingly becoming known. For example, ionizing radiation ( X- rays) and various toxic chemicals have been linked to mutations in animals and humans. Most mutations are spontaneous, however, and are simply DNA copying errors. The human mutation rate is higher in male sex cells (sperm) than in female sex cells (eggs) but is generally on the order of one per million per nucleotide per generation. The human genome includes about 3 billion base pairs, about 1.5% of which are protein- coding, so the average mutation rate in humans is .45 mutations

point mutations Replacements of a single nitrogen base with another base, which may or may not affect the amino acid for which the triplet codes.

synonymous point mutation A neutral point mutation in which the substituted nitrogen base creates a triplet coded to produce the same amino acid as that of the original triplet.

nonsynonymous point mutation A point mutation that creates a triplet coded to produce a different amino acid from that of the original triplet.

frameshift mutation The change in a gene due to the insertion or deletion of one or more nitrogen bases, which causes the subsequent triplets to be rearranged and the codons to be read incorrectly during translation.

transposable elements Mobile pieces of DNA that can copy themselves into entirely new areas of the chromosomes.

Klinefelter’s syndrome A chromosomal trisomy in which males have an extra X chromosome, resulting in an XXY con- dition; affected individuals typically have reduced fertility.

spontaneous mutations Random changes in DNA that occur during cell division.

induced mutations Refers to those muta- tions in the DNA resulting from exposure to toxic chemicals or to radiation.

mutagens Substances, such as toxins, chemicals, or radiation, that may induce genetic mutations.

T

A

Base substitution

(a) Base substitution

(b) Base insertion

A base substitution, the simplest kind of mutation, occurs when a single nitrogen base is substituted by another base. Here, thymine is being replaced by cytosine.

T

A

U

A

U

A

U

A

U

T

A

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

C

G

C

C

G

C

G

C

G

C

Valine Serine Serine Valine Proline

Base insertion

A

U

A

U

A

U

A

U

A

A

U

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

CG

C

G

C

Valine Serine Serine Tyrosine Serine

Normal gene

Mutation

Mutation

DNA template

mRNA

Protein

A

U

A

U

A

U

A

U

T

A

T

A

C

G

G

C

G

C

G

C

G

C

G

C

G

CG

C

G

C

Valine Serine Serine Isoleucine Proline

The mRNA receives the appropriate complementary nitrogen base, uracil. When the mRNA reaches the ribosomes, the series of codons is changed because of the insertion. Instead of reading the correct AUU, the ribosomes read UAU, which codes for the amino acid tyrosine instead of isoleucine. Again, this amino acid change may or may not affect the resulting protein.

A base insertion occurs when a nitrogen base is added to the DNA template. Here, adenine is inserted.

As a result of the base substitution, the mRNA strand receives complementary guanine rather than the adenine that pairs with thymine. This further substitution changes the amino acid that is added to the polypeptide chain at the ribosomes. Instead of isoleucine, valine is inserted (see Table 3.2 in chapter 3). This amino acid change may or may not affect the resulting protein.

FIGURE 4.5 DNA Mutations During the transcription phase of protein synthesis, errors in the DNA template can affect the resulting protein. Two types of DNA mutation are illustrated in these diagrams: (a) base substitutions and (b) base insertions.

 

 

80 | CHAPTER 4 Genes and Their Evolution: Population Genetics

in protein- coding genes per generation, or about one new, potentially significant mutation in every other person born.

For individuals, most mutations are relatively harmless, while a few may have profound consequences. For populations, mutations are inconsequential unless they offer selective adaptive advantages.

Natural Selection: Advantageous Characteristics, Survival, and Reproduction Darwin’s theory of evolution by means of natural selection provided the conceptual framework for understanding adaptation. That framework has become even more powerful over the last 150 years because it has allowed for many refinements. The principle of natural selection is based on Darwin’s conclusion that individuals with advantageous characteristics will survive in higher numbers and produce more

FIGURE 4.6 Spontaneous Mutations Spontaneous mutations can affect only physical appearance or can have health consequences, sometimes extreme ones. (a) A mutation called leucism has made this American alligator white. On its head are some spots of alligators’ normal, dark color. (b) The mutation that gives some cats white fur and blue eyes can also produce deafness and timidity. (c) Cheetahs normally have spotted coats, but a genetic mutation has produced stripes on this cheetah’s back. (d) Mutations can affect the wing count, eye color, and eye placement of fruit flies. A mutation has given this fruit fly abnormally placed, or ectopic, eyes, one of which is visible here as the red area on the wing.

(a)

(b)

(c) (d)

 

 

Natural Selection: Advantageous Characteristics, Survival, and Reproduction | 81

offspring than members of a population lacking advantageous characteristics. Natural selection, therefore, focuses on reproductive success, or fitness. In par- ticular, fitness is defined as some measure of the propensity to contribute offspring to future generations, usually by the next generation. Fitness can be defined in reference to individuals in a population or to specific genotypes. For our purposes, fitness is defined on the basis of genotypes. This means that some genotypes have more (or less) fitness than other genotypes. The implication is that fitness differ- ences can result in changes to allele frequencies. For example, if the genotype for darker coloring confers an adaptive advantage over the genotype for lighter color- ing, the dark- color genotype will likely increase in frequency over time.

PATTERNS OF NATURAL SELECTION Evolutionary biologists have identified three alternative patterns by which natural selection can act on a specific trait. Directional selection favors one extreme form of a trait— more children are produced by individuals who have that extreme trait, so selection moves in that direction. Human evolution, for example, has clearly favored larger brains (for more on these topics, see chapters 10 and 11). Stabilizing selection favors the average version of a trait. For example, living humans whose birth weights are in the middle of the range have a better chance of surviving and reproducing than do those born with the lowest and highest weights. In disruptive selection, the pattern of variation is discontinuous. Individuals at both extreme ends of the range produce more offspring than do the remainder of the population. Given enough time, this pattern can result in a speciation event as those in the middle fail to survive and reproduce and two new species arise at the extremes (Figure 4.7).

fitness Average number of offspring produced by parents with a particular genotype compared to the number of off- spring produced by parents with another genotype.

directional selection Selection for one allele over the other alleles, causing the allele frequencies to shift in one direction.

stabilizing selection Selection against the extremes of the phenotypic distribution, decreasing the genetic diversity for this trait in the population.

disruptive selection Selection for both extremes of the phenotypic distribution; may eventually lead to a speciation event.

Fr eq

ue nc

y

Body size

Directional selection

Body size

Stabilizing selection

Body size

Disruptive selection

Body size

Fi tn

es s

(n um

be r o

f of

fs pr

in g

pr od

uc ed

)

Body size Body size Body size Body size

Av er

ag e

si ze

(i n

po pu

la tio

n)

Time Time Time Time

No selection

(a) (b) (c) (d) FIGURE 4.7 Types of Selection (a) Top: In the population represented here, smaller body size is more favorable than larger body size, so the frequency of smaller body size will increase thanks to directional selection. Middle: The fitness of individuals with smaller body sizes will be greater than that of individuals with larger body sizes. Bottom: Over time, the population’s average body size will decrease. (b) Top: In this population, medium body size is favored, so the frequency of medium body size will increase thanks to stabilizing selection. Middle: The fitness of individuals with medium body sizes will be much greater. Bottom: However, the population’s average body size will remain relatively stable over time. (c) Top: Here, owing to disruptive selection, the frequencies of small and large body sizes will increase, while the frequency of medium body size will decrease. Middle: The fitness levels are highest at the extremes and lowest in the middle. Bottom: Over time, the population will split between those with large bodies and those with small bodies. (d) Top: In the absence of selection, the population will have a range of sizes. Middle: Fitness levels will vary independently of size. Bottom: The population’s average body size will not change over time.

 

 

82 | CHAPTER 4 Genes and Their Evolution: Population Genetics

NATURAL SELECTION IN ANIMALS: THE CASE OF THE PEPPERED MOTH AND INDUSTRIAL MELANISM Examples of natural selection in animals are wide- ranging. Among them are animals that “blend in” with their surrounding habitat (Figure 4.8). Perhaps the best evidence ever documented of natural selection operating on a heritable trait concerns the peppered moth, Biston betularia, a species common throughout Great Britain (Figure  4.9). This moth is nocturnal, eating and breeding by night and attaching itself to trees, especially in the upper branches, during the day. Prior to the mid- 1800s, all members of the species had a peppered appearance, their white coloring sprinkled with black. Trees throughout Great Britain were covered with lichen, and the moths’ coloration provided excellent camouflage against the trees’ variable- colored surface and thus protected the moths from their major predator, birds. In 1848, a naturalist exploring the countryside near Manchester, England, spotted a completely black variety of the moth. A new species name, Biston car- bonaria, distinguished this melanic (dark) form from the nonmelanic (light) form. The frequency of the melanic form remained relatively low for a couple of decades but climbed rapidly in the late nineteenth century. By the 1950s, 90% or more of peppered moths were melanic (Figure 4.10).

This rapid increase in melanic frequency was a case of evolution profoundly changing phenotype. Directional selection had favored the melanic form over the nonmelanic form, and the melanic form exhibited a greater fitness. But what was this form’s adaptive advantage?

The selecting factor was the Industrial Revolution. With the rise of industry throughout England and elsewhere in the middle to late nineteenth century, mills, fueled entirely by coal, spewed coal particles from smokestacks— 50 tons per square mile per month, in some places— blackening the sky and covering the landscape. The trees survived this pollution onslaught, but the lichen covering the trees did not. The trees’ surfaces went from light- colored to black, greatly altering the peppered moth’s habitat. This pollution crisis provided a huge selective advantage for the melanic moths, which were now perfectly camouflaged against blackened trees. Nonmelanic moths became easy prey.

FIGURE 4.8 Leafy Sea Dragon As a result of natural selection, the leafy sea dragon looks just like the sea plants around it in its ocean setting.

melanic Refers to an individual with high concentrations of melanin.

nonmelanic Refers to an individual with low concentrations of melanin.

FIGURE 4.9 Peppered Moths The genus Biston includes two species: Biston betularia (light) and Biston carbonaria (dark).

 

 

Natural Selection: Advantageous Characteristics, Survival, and Reproduction | 83

How did the genetics of this evolutionary change work? Breeding experiments revealed that, in a classic case of Mendelian genetics, the color difference between the two Biston species was determined by one locus. The nonmelanic variety had a genotype of cc (homozygous recessive), while the melanic variety was either het- erozygous, Cc, or homozygous,  CC.  The dominant allele, C, likely first appeared as a mutation, perhaps in the first half of the nineteenth century or earlier, long before the first melanic moth was observed, in 1848. The C allele may have been in the population, maintained by the mutation’s reoccurrence. Recent estimates suggest that the frequency of the nonmelanic variety was only 1%–10% in polluted regions of England and no more than 5% around Manchester. Plugged into the Hardy- Weinberg equilibrium, this information in turn suggests that 46% of the population had the CC genotype, 44% had the Cc genotype, and 10% had the cc genotype (Table 4.2).

TABLE 4.2 Moth Genotype Frequencies— Industrialization Period

Phenotype Melanic Nonmelanic

Phenotype Frequencies .90 .10

Genotype CC + Cc cc

Genotype Frequencies p2 + 2pq q2

Allele Frequency* Calculations:

Step 1 q2 = .10 (q2 = f[cc])

Step 2 q = √q2 = √.10 = .32

Step 3 p = 1 − q = 1 − .32 = .68

Genotype Frequency Calculations: p2 = f(CC) = .682 = .46

2pq = f(Cc) = 2 × .68 × .32 = .44

q2 = f(cc) = .322 = .10

Check: p2 + 2pq + q2 = 1

.46 + .44 + .10 = 1

*f = frequency

FIGURE 4.10 Changes in the Peppered Moth Gene Frequency The frequency of the melanic gene in peppered moths increased from 1848 through 1948. The frequency of the nonmelanic gene decreased during that same period.

G en

e Fr

eq ue

nc y

Generation Date

1.0

0.8

0.1

0.2

0.9

0.3

0.4

0.5

0.6

0.7

1848 1858 1868 1878 1888 1898 1908 1918 1928 1938 1948 0

 

 

84 | CHAPTER 4 Genes and Their Evolution: Population Genetics

Beginning in the late 1960s and early 1970s, the stricter pollution laws, changes in coal burning, and decline of mill- based industry in Great Britain profoundly affected the moth population, once again illustrating natural selection. That is, in areas that were no longer polluted, the frequency of Biston carbonaria dropped. In Manchester, for example, the frequency of the melanic moth decreased from 90% in 1983 to well under 10% in the late 1990s. Plugged into the Hardy- Weinberg equilibrium, these numbers reveal that the CC genotype decreased to 0.25%, the Cc genotype decreased to 9.5%, and the cc genotype increased to 90% (Table 4.3). This rapid evolutionary change reflected the return of the trees’ original col- oration, which conferred a selective disadvantage— predation visibility— on the melanic variety. This postscript adds even more power to the story of how natural selection brought about biological changes in the genus Biston.

NATURAL SELECTION IN HUMANS: ABNORMAL HEMOGLOBINS AND RESISTANCE TO MALARIA The above case of industrial melanism is an example of positive selection, whereby an organism’s biology is shaped by selection for beneficial traits. Natural selection for beneficial traits in humans is best understood by studying genes that control specific traits. Of the 90 or so different loci that are targets of natural selection (Figure  4.11), among the most compelling examples is the sickle- cell allele— the hemoglobin S (or simply S) allele— which causes sickle- cell anemia (Figure  4.12). Millions of people suffer from such hemolytic anemias, which involve the destruction of red blood cells. A low number of red blood cells can produce health problems because of the resultant lack of hemoglobin, the chemical in red blood cells that carries oxygen to all the body tissues. The S gene yields a specific kind of abnormal hemoglobin.

positive selection Process in which advan- tageous genetic variants quickly increase in frequency in a population.

sickle- cell anemia A genetic blood dis- ease in which the red blood cells become deformed and sickle- shaped, decreasing their ability to carry oxygen to tissues.

hemolytic anemias Conditions of insuf- ficient iron in the blood due to the destruction of red blood cells resulting from genetic blood diseases, toxins, or infectious pathogens.

abnormal hemoglobin Hemoglobin altered so that it is less efficient in binding to and carrying oxygen.

TABLE 4.3 Moth Genotype Frequencies— Postindustrialization Period

Phenotype Melanic Nonmelanic

Phenotype Frequencies .10 .90

Genotype CC + Cc cc

Genotype Frequencies p2 + 2pq q2

Allele Frequency* Calculations:

Step 1 q2 = .90 (q2 = f[cc])

Step 2 q = √q2 = √.90 = .95

Step 3 p = 1 − q = 1 − .95 = .05

Genotype Frequency Calculations: p2 = f(CC) = .052 = .0025

2pq = f(Cc) = 2 × .05 × .95 = .095

q2 = f(cc) = .952 = .90

Check: p2 + 2pq + q2 = 1

.0025 + .095 + .90 = 1

*f = frequency

 

 

Natural Selection: Advantageous Characteristics, Survival, and Reproduction | 85

Sickle- cell anemia has been known since the early 1900s, and the genetics behind it were documented in the 1950s. The S gene is a simple base- pair mutation (Figure  4.13). Genetically, people with normal hemoglobin have the alleles AA, the homozygous condition. People who carry the sickle- cell allele on one of the two homologous chromosomes only are AS, and people who have the homozygous form of the disease are SS. AS individuals are for all practical purposes normal in their survival and reproduction rates. There is no cure for sickle- cell anemia, and in the absence of modern medical treatment, some 80% of people who are SS die before the reproductive years, usually considerably earlier. The SS genotype results in many red blood cells’ having a sickle shape caused by the abnormal hemoglobin, in sharp contrast to the round appearance of red blood cells in people with normal hemoglobin (Figure 4.14). The cells’ abnormal shape prevents them from passing through the capillaries, the narrow blood vessels that form networks throughout tissues. When the clogging of capillaries cuts off the oxygen supply in vital tissues, severe anemia and death can result.

Chromosome 11

Hemoglobin S

X chromosome

G6pd

FIGURE 4.11 G6pd One target of natural selection in humans is the G6pd gene, located on the X chromosome.

FIGURE 4.12 Sickle- Cell Gene Hemoglobin S appears on human chromosome 11.

capillaries Small blood vessels between the terminal ends of arteries and the veins.

 

 

86 | CHAPTER 4 Genes and Their Evolution: Population Genetics

THE GEOGRAPHY OF SICKLE- CELL ANEMIA AND THE ASSOCIATION WITH MALARIA Beginning in the mid- twentieth century, the medical commu- nity observed that many people living in equatorial Africa— as many as 20%– 30%—had the S gene. This finding represented a huge puzzle: since the gene was so bad for survival, why was its frequency so high? In other words, one would expect strong selection against this nonbeneficial gene. The solution to the puzzle began to emerge with the discovery that high heterozygous (AS) frequencies appear in regions of Africa where malaria is endemic. In other words, where malaria— a potentially lethal parasitic infection in which the parasite is introduced to a human host by a mosquito— is always present, there is a high frequency of carriers of the gene (Figure 4.15). Moreover, AS people ( sickle- gene carriers) die of malaria in far fewer numbers than do AA people.

As described at the beginning of this chapter, Anthony Allison discovered that in low- lying, wet areas of Kenya (where the number of mosquitoes was great and the rate of malaria was high), the frequency of the sickle- cell allele was consid- erably higher than in highland or arid areas. He developed the simple but elegant hypothesis that the infection and the genetic mutation were related. Individuals homozygous for normal hemoglobin (AA) were highly susceptible to dying from malaria; individuals homozygous for sickle- cell anemia (SS) did not survive to reproduce; however, individuals heterozygous for normal hemoglobin and the sickle- cell mutation (AS) either did not contract malaria or suffered a less severe malarial infection. That these frequencies were being maintained indicated that the AS heterozygote was a balanced polymorphism. It was also a fitness trade- off: car- riers could pass on the sickle- cell allele, but they received immunity from malaria.

FIGURE 4.14 Sickle- Shaped Red Blood Cells This image shows normal red blood cells, which are round; a long, slender, sickle- shaped cell (at top); and other irregularly shaped cells. The abnormal cells are very fragile and easily damaged or destroyed.

The S allele arises from a nitrogen base substitution: adenine replaces thymine in the hemoglobin DNA.

The codon changes from GAG in normal hemoglobin mRNA to GUG in the sickle-cell hemoglobin mRNA (the adenine in the normal mRNA is replaced with uracil).

A

U

T

A

C

G

C

G

Transcription

Normal hemoglobin DNA

C

G

C

G

Sickle-cell hemoglobin DNA

mRNA

Translation

Normal hemoglobin

Sickle-cell hemoglobin

mRNA

The changed codon in the mRNA is coded to produce a different amino acid. Instead of glutamate, valine is inserted in the polypeptide chain. This amino acid change causes a change in the functioning of the hemoglobin so that it cannot bind to oxygen.

Glutamatic acid Valine

FIGURE 4.13 Sickle- Cell Mutation Sickle- cell anemia begins with a single nitrogen base mutation— a base substitution. The abnormal hemoglobin that results is less efficient at binding oxygen and causes red blood cells to become sickle- shaped.

 

 

Natural Selection: Advantageous Characteristics, Survival, and Reproduction | 87

THE BIOLOGY OF SICKLE- CELL ANEMIA AND MALARIAL INFECTION Why do people who are heterozygous for the sickle- cell allele survive malaria or not con- tract it at all? Unlike SS red blood cells, AS red blood cells do not sickle under most conditions (that is, except when severely deprived of oxygen). They are, however, somewhat smaller than normal cells, and their oxygen levels are somewhat lower. For reasons not yet understood, the AS red blood cells are simply a poor host— a nonconducive living and reproduction environment— for the parasite that causes malaria (Figure 4.16).

THE HISTORY OF SICKLE- CELL ANEMIA AND MALARIA In the late 1950s, the American physical anthropologist Frank B. Livingstone (1928–2005) sought to strengthen the case for natural selection by historically linking sickle- cell anemia and malaria. Livingstone asked two important questions: Where and when did the sickle- cell allele first appear in equatorial Africa? and What conditions led to the allele’s being naturally selected? He hypothesized that the Bantu, a group of peoples who speak Bantu languages, carried the mutation with them when they migrated south- ward from the region of Cameroon and Nigeria (Figure 4.17). Prior to the Bantu’s arrival, the region was a largely unbroken forest. Bantu populations introduced agri- culture there, clearing large swaths of the forest for cultivation. The peoples’ iron- working technology made possible the creation of tools for cutting down large trees, clearing and plowing fields, and planting crops— mostly yams and cassava.

Even under the best conditions, tropical forests are fragile ecosystems. Once their trees have been cleared and their fields have been planted, their relatively poor soil, which normally soaks up rainwater, becomes thin or disappears. Geo- logical evidence shows a dramatic increase in soil erosion in the region after the arrival of Bantu populations, due in large part to anthropogenic deforestation and the overall environmental impact of humans on the landscape. As a result of these erosive processes, pools of water collect and become stagnant, providing ideal conditions for the breeding of parasite- carrying mosquitoes (Figure  4.18). This picture became clear to Livingstone as he developed his research: the newly created ecological circumstances fostered mosquito reproduction and the spread of malaria, and the growing host of humans made possible by agriculture- fueled population growth provided the food resources needed by the mosquitoes. The

(b)(a)

Greater than 14% 12–14% 10–11.9% 8–9.9% 6–7.9% 4–5.9% 2–3.9% 0–1.9%

Areas where malaria is present

FIGURE 4.15 Distributions of the Sickle- Cell Allele and Malaria In equatorial Africa, (a) the distribution of the sickle- cell allele coincides with (b) areas of high malarial parasite concentration.

balanced polymorphism Situation in which selection maintains two or more phenotypes for a specific gene in a population.

anthropogenic Refers to any effect caused by humans.

 

 

88 | CHAPTER 4 Genes and Their Evolution: Population Genetics

infectious disease gave those individuals with a very rare mutation— the sickle- cell allele— an adaptive advantage and the ability to survive and reproduce in these new environmental circumstances. Due to the advantage the heterozygous condition provides, the S allele was maintained and passed from generation to generation. For this reason, sickle- cell anemia predominantly affects those whose descendants came from the malarial environments in large parts of equatorial Africa. Outside of such malarial environments, the S allele never became advantageous.

OTHER HEMOGLOBIN AND ENZYME ABNORMALITIES  Sickle- cell anemia turns out to be just one of a number of hemoglobinopathies and other genetic abnormalities in Africa, Asia, and Europe that provide a strong selective advantage in regions of endemic malaria (Figure  4.19). Heterozygous carriers of abnormal hemoglobins apparently make poor hosts for malarial parasites.

Thalassemia, a genetic anemia found in Europe (especially in Italy and Greece), Asia, and the Pacific, reduces or eliminates hemoglobin synthesis. In some homozygous forms of the mutation, hemoglobin becomes clumped inside the red blood cells. The spleen then destroys the red blood cells, resulting in severe anemia. In the areas around the Mediterranean where the genetic frequency is highest— as high as 80%—the presence of malaria makes a strong case for a selective advantage for heterozygous individuals, for whom the condition and malaria are not lethal.

An association has long been recognized between deficiency of the enzyme glucose- 6-phosphate dehydrogenase (G6PD) and malaria. A recessive hered- itary mutation leads more males than females to lack the gene that is coded to produce this enzyme (see Figure 4.11). Without the G6PD enzyme, a person who takes sulfa- based antibiotics or eats fava beans risks the destruction of red blood cells, severe anemia, and occasionally death. Because of the connection with fava beans, this severe hemolytic disease is called favism. Its 130 genetic variants occur in high frequencies in some populations, the highest being 70% among Kurdish Jews. Heterozygous carriers have a strong selective advantage because they pro- duce some of the enzyme but are protected from malaria (here again, the parasite cannot live in the abnormal red blood cells).

Analysis of genetic data by the anthropologist Sara Tishkoff indicates that the mutation for the disease arose between about 4,000 and 12,000 yBP, at the same time as the abnormal hemoglobins. Populations whose descendants did not encounter malaria do not have the G6pd mutation or abnormal hemoglobins.

Once inside the human, the sporozoites travel through the bloodstream to the liver. In the liver, the sporozoites create thousands of merozoites. A merozoite is a daughter cell that results from asexual reproduction.

The newly produced merozoites enter the bloodstream and infect the red blood cells. Within these cells, the merozoites continue to multiply, eventually causing the cells to rupture. The merozoites then invade other red blood cells in the bloodstream, and the cycle continues.

a Sporozoite: The goal of sporozoite vaccines is to block parasites from entering or growing within human liver cells.

b Merozoite: Vaccines based on merozoite antigens lessen malaria’s severity by hobbling the invasion of new generations of red blood cells or by reducing complications.

c Gametocyte: So-called altruistic gametocyte- based vaccines do not affect human disease but are designed to evoke human antibodies that derail parasite development within the mosquito.

Over time, some of the merozoites develop into male and female gametocytes, which may be transferred to another mosquito that bites the human host. Gametocytes are cells that can divide to produce gametes, or sex cells.

Within the mosquito, sexual reproduction occurs, as the male gamete fertilizes the female gamete. Oocysts, or zygotes, develop and even- tually release sporozoites. These newly formed sporozoites travel to the salivary glands of the mosquito, where they can be transmitted to the next human the mosquito bites.

Within the mosquito, sexual reproduction occurs, as the male gamete fertilizes the female gamete. Oocysts, or zygotes, develop and even- tually release sporozoites. These newly formed sporozoites travel to the salivary glands of the mosquito, where they can be transmitted to the next human the mosquito bites.

Merozoites

Host’s liver

Host’s red blood cell

INSIDE HUMAN

VACCINE TARGETS

INSIDE MOSQUITO

Female gametocyte

Female gamete

Male gamete

Oocyst

Male gametocyte

Fertilizationc

Sporozoites

A mosquito bites a human, passing sporozoites to the new host. A sporozoite is a motile form of the parasite.

a

b

1

5 2

4

3

FIGURE 4.16 The Spread of Malaria The life cycle of the malarial parasite, Plasmodium falciparum, takes place in two hosts: mosquito and human. Both hosts are needed if the parasite is to survive.

 

 

Natural Selection: Advantageous Characteristics, Survival, and Reproduction | 89

infectious disease gave those individuals with a very rare mutation— the sickle- cell allele— an adaptive advantage and the ability to survive and reproduce in these new environmental circumstances. Due to the advantage the heterozygous condition provides, the S allele was maintained and passed from generation to generation. For this reason, sickle- cell anemia predominantly affects those whose descendants came from the malarial environments in large parts of equatorial Africa. Outside of such malarial environments, the S allele never became advantageous.

OTHER HEMOGLOBIN AND ENZYME ABNORMALITIES  Sickle- cell anemia turns out to be just one of a number of hemoglobinopathies and other genetic abnormalities in Africa, Asia, and Europe that provide a strong selective advantage in regions of endemic malaria (Figure  4.19). Heterozygous carriers of abnormal hemoglobins apparently make poor hosts for malarial parasites.

Thalassemia, a genetic anemia found in Europe (especially in Italy and Greece), Asia, and the Pacific, reduces or eliminates hemoglobin synthesis. In some homozygous forms of the mutation, hemoglobin becomes clumped inside the red blood cells. The spleen then destroys the red blood cells, resulting in severe anemia. In the areas around the Mediterranean where the genetic frequency is highest— as high as 80%—the presence of malaria makes a strong case for a selective advantage for heterozygous individuals, for whom the condition and malaria are not lethal.

An association has long been recognized between deficiency of the enzyme glucose- 6-phosphate dehydrogenase (G6PD) and malaria. A recessive hered- itary mutation leads more males than females to lack the gene that is coded to produce this enzyme (see Figure 4.11). Without the G6PD enzyme, a person who takes sulfa- based antibiotics or eats fava beans risks the destruction of red blood cells, severe anemia, and occasionally death. Because of the connection with fava beans, this severe hemolytic disease is called favism. Its 130 genetic variants occur in high frequencies in some populations, the highest being 70% among Kurdish Jews. Heterozygous carriers have a strong selective advantage because they pro- duce some of the enzyme but are protected from malaria (here again, the parasite cannot live in the abnormal red blood cells).

Analysis of genetic data by the anthropologist Sara Tishkoff indicates that the mutation for the disease arose between about 4,000 and 12,000 yBP, at the same time as the abnormal hemoglobins. Populations whose descendants did not encounter malaria do not have the G6pd mutation or abnormal hemoglobins.

Once inside the human, the sporozoites travel through the bloodstream to the liver. In the liver, the sporozoites create thousands of merozoites. A merozoite is a daughter cell that results from asexual reproduction.

The newly produced merozoites enter the bloodstream and infect the red blood cells. Within these cells, the merozoites continue to multiply, eventually causing the cells to rupture. The merozoites then invade other red blood cells in the bloodstream, and the cycle continues.

a Sporozoite: The goal of sporozoite vaccines is to block parasites from entering or growing within human liver cells.

b Merozoite: Vaccines based on merozoite antigens lessen malaria’s severity by hobbling the invasion of new generations of red blood cells or by reducing complications.

c Gametocyte: So-called altruistic gametocyte- based vaccines do not affect human disease but are designed to evoke human antibodies that derail parasite development within the mosquito.

Over time, some of the merozoites develop into male and female gametocytes, which may be transferred to another mosquito that bites the human host. Gametocytes are cells that can divide to produce gametes, or sex cells.

Within the mosquito, sexual reproduction occurs, as the male gamete fertilizes the female gamete. Oocysts, or zygotes, develop and even- tually release sporozoites. These newly formed sporozoites travel to the salivary glands of the mosquito, where they can be transmitted to the next human the mosquito bites.

Within the mosquito, sexual reproduction occurs, as the male gamete fertilizes the female gamete. Oocysts, or zygotes, develop and even- tually release sporozoites. These newly formed sporozoites travel to the salivary glands of the mosquito, where they can be transmitted to the next human the mosquito bites.

Merozoites

Host’s liver

Host’s red blood cell

INSIDE HUMAN

VACCINE TARGETS

INSIDE MOSQUITO

Female gametocyte

Female gamete

Male gamete

Oocyst

Male gametocyte

Fertilizationc

Sporozoites

A mosquito bites a human, passing sporozoites to the new host. A sporozoite is a motile form of the parasite.

a

b

1

5 2

4

3

FIGURE 4.16 The Spread of Malaria The life cycle of the malarial parasite, Plasmodium falciparum, takes place in two hosts: mosquito and human. Both hosts are needed if the parasite is to survive.

FIGURE 4.17 Bantu Expansion Beginning by about 1000 BC, Bantu people began a series of migrations, originating in central Africa and pushing southward eventually into southern Africa. In addition to agriculture, they carried the mutation for the sickle- cell allele.

Bantu Khoisan

hemoglobinopathies A group of related genetic blood diseases characterized by abnormal hemoglobin.

thalassemia A genetic blood disease in which the hemoglobin is improperly synthesized, causing the red blood cells to have a much shorter life span.

glucose- 6-phosphate dehydrogenase (G6PD) An enzyme that aids in the proper functioning of red blood cells; its deficiency, a genetic condition, leads to hemolytic anemia.

(a) (b)

FIGURE 4.18 The Spread of Malaria As the Bantu cleared forests for agricultural fields, (a) pools of stagnant water (b) became an ideal breeding ground for mosquito larvae, which carried the malarial parasites.

 

 

90 | CHAPTER 4 Genes and Their Evolution: Population Genetics

Today, for example, malaria appears throughout the tropical regions of the Amer- icas, but Native Americans are 100% homozygous for normal alleles at the G6pd and hemoglobin loci. That these particular genes do not appear to have mutated in the New World strongly suggests that malaria was introduced to North and South America only after the Europeans’ arrival. Indeed, the introduction of malaria and other Old World diseases— by either Spaniards or their African slaves— likely played an instrumental role in the precipitous decline in the native populations.

If malaria had been introduced in the Americas much earlier than the last few centuries— say thousands of years ago— and the mutations occurred, there might have been time for a natural selection to develop for the mutations. If the mutations did appear before the Europeans’ arrival, however, they would have exhibited a clear selective disadvantage in the absence of malaria and been weeded out of the gene pool. Thus, red blood cell polymorphisms in the abnormal hemoglobin and G6pd loci reflect the fundamental interactions among environment, genes, and culture that have resulted in the modern human genome. The genes provide an important record about human evolution and the role of natural selection in shaping genetic variation.

Genetic Drift: Genetic Change due to Chance One of the four forces of evolution (see “The Evolutionary Synthesis, the Study of Populations, and the Causes of Evolution” in chapter  2), genetic drift, is random change in allele frequency over time. Provided that no allele confers a selective advantage over another, a random change can lead to a change in gene frequency, such as one allele being lost and the other becoming fixated— or fixed, the only allele of its kind, in the population. This force, this kind of change, makes possible the measuring of evolution as a statistical probability.

Hb C Hb E Hb S

FIGURE 4.19 Distribution of Hemoglobinopathies This map shows the distribution of hemoglobin E in Southeast Asia. People with hemoglobin E, an uncommon but severe blood abnormality, may have mild hemolytic anemia or other mild effects. Like hemoglobin S, hemoglobin C appears primarily in equatorial Africa. Like hemoglobin E, hemoglobin C has generally minor effects, most often mild hemolytic anemia.

 

 

Genetic Drift: Genetic Change due to Chance | 91

Coin tosses can demonstrate the effects of genetic drift (Table  4.4). Imagine that heads and tails are two alleles in a population. If there are only two members of the population (two coin tosses), there is a great chance that both will be heads. In effect, the “heads” allele will become fixed in the small population, while the “tails” allele will be lost. As the population size (number of coin tosses) increases, it becomes less likely that one allele will become fixed and the other lost. In very large populations (1  million coin tosses), both alleles may be present in equal proportions.

How does such statistical probability translate to populations? Now imagine that before the election of your student government you have been asked to predict the winners. The best way for you to predict would be to ask each voting- eligible student how he or she planned to vote. It is highly likely that the outcome of such a comprehensive poll would be close to the actual election results. The shortcom- ings of this approach might include the very large size of the target population. No one would interview, for example, 50,000 students! The second- best approach would be to select a sample, preferably a random sample that represented the entire student body. If you selected five students out of the 50,000, chances are very slim that those five would represent all the ethnic, national, regional, and economic backgrounds of the student body. In fact, chances are very high that this sample (.01% of the total population) would provide a voting outcome very different from the actual one. If you interviewed 500 students (1% of the population), the chances of representation would be much greater; and if you chose 5,000 students (10% of the population), they would be greater still. The larger the sample size, the greater the probability of an accurate prediction.

Variations in human populations work the exact same way, except that genetic drift operates over a period of time rather than at a single point. The probability of an allele’s frequency changing in a relatively short period of time increases with

TABLE 4.4 Heads versus Tails: Genetic Drift and Probability

Coin Tosses Heads Tails Heads: Tails Rat io

2 2 0 2:0

10 4 6 4:6

50 22 28 11:14

100 55 45 11:9

200 199 201 199:201

500 253 247 253:247

1,000 501 499 501:499

5,000 2,500 2,500 1:1

10,000 5,000 5,000 1:1

100,000 50,000 50,000 1:1

500,000 250,000 250,000 1:1

1,000,000 500,000 500,000 1:1

 

 

92 | CHAPTER 4 Genes and Their Evolution: Population Genetics

decreasing population size. The larger the population, the less divergence from the original gene frequency over time (Figure 4.20).

How does this effect play out in real life? Among humans, for example, genetic drift might occur in a small group that is endogamous, discouraging reproduc- tion outside the group. (An exogamous society extends reproduction outside its community.) Within such a group, the chances are great that the frequencies of genetic markers will differ from those of a larger population. When the Dunkers, a small religious sect that discourages outside marriage (and thus reproduction), first emigrated from Germany to Pennsylvania, in 1719, the group included just 28 members. Over the next few decades, several hundred more arrived in Pennsylvania; the breeding population remained quite small. Comparisons of contemporary blood type percentages among Dunkers, Germans, and Americans reflect significant changes in the Dunkers and a likely lack of change in the larger populations (Figure  4.21). That is, blood type frequencies among Germans and Americans remain basically the same as they were in the 1700s. The Dunkers’ original frequencies were probably much like those of the Germans, but the small

endogamous Refers to a population in which individuals breed only with other members of the population.

exogamous Refers to a population in which individuals breed only with non- members of their population.

founder effect The accumulation of random genetic changes in a small population that has become isolated from the parent population due to the genetic input of only a few colonizers.

G en

e fre

qu en

cy (%

)

Number of generations

100

50

Found something interesting ?

• On-time delivery guarantee
• PhD-level professional writers
• Free Plagiarism Report

• 100% money-back guarantee
• Absolute Privacy & Confidentiality
• High Quality custom-written papers

Related Model Questions

Feel free to peruse our college and university model questions. If any our our assignment tasks interests you, click to place your order. Every paper is written by our professional essay writers from scratch to avoid plagiarism. We guarantee highest quality of work besides delivering your paper on time.

Grab your Discount!

25% Coupon Code: SAVE25
get 25% !!