Red Desert: History of a Place 9780292786059

A vast expanse of rock formations, sand dunes, and sagebrush in central and southwest Wyoming, the little-known Red Dese

204 82 52MB

English Pages 412 Year 2012

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Red Desert: History of a Place
 9780292786059

Citation preview

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page i

red desert

Proulx_00_pp.00i-xii:Layout 1

8/21/08

1:04 PM

Page ii

Willow Hill and old Rawlins-Bairoil Road, Carbon County, Wyoming, 2004

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page iii

Red Desert

history of a place

Edited by annie proulx Photographs by martin stupich

university of texas press

austin

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page iv

The publication of this book was made possible by a generous contribution from the University of Texas Press Advisory Council.

Copyright © 2008 by Dead Line Ltd. Photographs copyright © 2008 by Martin Stupich All rights reserved Printed in Singapore First edition, 2008 Requests for permission to reproduce material from this work should be sent to: Permissions university of texas press P.O. Box 7819 Austin, TX 78713-7819 www.utexas.edu/utpress/about/bpermission.html The paper used in this book meets the minimum requirements of ansi/niso z39.48-1992 (r1997) (Permanence of Paper). Library of Congress Cataloging-in-Publication Data Red Desert : history of a place / edited by Annie Proulx ; photographs by Martin Stupich. — 1st ed. p. cm. isbn 978-0-292-71420-5 (cl. : alk. paper) Includes bibliographical references and index. 1. Natural history—Wyoming—Red Desert (Desert). 2. Red Desert (Wyo. : Desert)—History. I. Proulx, Annie, date II. Stupich, Martin, date qh105.w8 r43 2008 508.787— dc22 2008011077

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page v

This book is for all who love the cranky, sagebrushed, rare, and fragile Red Desert of Wyoming. —Annie Proulx For David Harriet and John —Martin Stupich

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page vi

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page vii

contents

List of Illustrations Abbreviations x

ix

I. PHOTOGRAPHS Acknowledgments 3 Photographs 4 II. TEXT Acknowledgments 75 Introduction 77 Natural History 1. Geology of the Red Desert, by Charles Ferguson 85 2. Water in the Red, by Craig Thompson 107 3. Environmental Change in the Wyoming Basin’s Red Desert, by Dudley Gardner 121 4. Titanotheres, Time, and People: A Snapshot of Red Desert Paleontology, by Tom Rea 137 5. Vertebrate Wildlife of the Red Desert, by Gary P. Beauvais 143 6. Birds of the Red Desert, by Andrea Orabona 171 7. Horses Come to the Red Desert, by Dudley Gardner 181 8. Insects of the Red Desert: An Exercise in Scientific Humility, by Jeffrey A. Lockwood 189 9. Sagebrush, by George P. Jones 201 10. Bright Green Hues Are Rare: Plant Diversity and Conservation in Wyoming’s Red Desert, by Walter and Laura Fertig 217 11. Biological (Cryptobiotic) Soil Crusts of the Red Desert, by Jack States 227

Human History 12. Early People of the Red Desert, by Dudley Gardner 231 13. The Shoshonis and Westward-Bound Emigrants, by Dudley Gardner 239 14. An Anthropological Impression of Rock Art in the Greater Red Desert, by Russel L. Tanner 245 15. Traversing the Desert, by Annie Proulx 253 16. Forts of the Red Desert, by Annie Proulx 267 17. Fort Bridger and Camps Stambaugh and Pilot Butte, by Dudley Gardner 271 18. Forts Halleck and Fred Steele, by Annie Proulx 283 19. The Union Pacific Railroad Arrives, by Annie Proulx 293 20. The Union Pacific, the Chinese, and the Japanese, by Dudley Gardner 297 21. Inhabitants of the Margins, by Annie Proulx 305 22. The Little Snake River Valley, by Annie Proulx 311 23. Red Desert Ranches, by Annie Proulx 317 24. Horse Bands of the Red Desert, by Annie Proulx 329 25. Opening the Oyster, by Annie Proulx 339 26. Red Desert Outlaws, by Annie Proulx 355 27. History of Conservation Efforts in the Red Desert, by Mac Blewer 363 Contributors Index 377

375

vii

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page viii

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page ix

illustrations

MAPS

TABLES

1. The Greater Red Desert endpapers

1.1. Geologic time—the eons, eras, and periods of earth

2. Geologic map of the Red Desert and its environs 87

history and the epochs of the Cenozoic

86

3. Water features of the Red Desert

112

3.1. Generalized climate regions from 20,000 to 150 bp

4. Range of white-tailed prairie dog

149

5.1. Examples of peripheral vertebrates occupying the

5. Range of Wyoming pocket gopher 6. Range of Wyoming ground squirrel

150

Red Desert

151

8. Camps and forts in the Red Desert

Red Desert 289

148

5.4.Vertebrates that are endemic to the Red Desert 6.1.Bird species mentioned in Chapter 6

FIGURES

148

173

9.1 Sagebrush species and subspecies and their occurrence

5.1. White-tailed prairie dog 5.2. Wyoming pocket gopher

149

in the Red Desert

202

150

5.3. Wyoming ground squirrel 8.1. Malaise trap

145

5.3. Examples of widespread vertebrates occupying the

268

9. Thornburgh’s route to the White River Agency

144

5.2.Wildlife species mentioned in Chapter 5

7. Cherokee Trail routes, 1849 –1850 257

122

151

191

8.2.Insects drawn to a light trap

192

ix

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page x

abbreviations

bbs

Breeding Bird Survey

blm

Bureau of Land Management

bp

before present

bsc

biological soil crust

cbm

coal-bed methane

cwa

Civil Works Administration

hma

herd management area

ma

millions of years

shpo

State Historic Preservation Office (Wyoming)

usda

U.S. Department of Agriculture

usdi

U.S. Department of the Interior

usfs

U.S. Forest Service

usgs

U.S. Geological Survey

wpa

Works Progress Administration

wwcc Western Wyoming Community College

x

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page xi

red desert

Proulx_00_pp.00i-xii:Layout 1

8/4/08

6:03 PM

Page xii

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 1

PART I photographs

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 2

Proulx_01_pp.001-039:Layout 1

8/7/08

4:05 PM

Page 3

acknowledgments

Dozens of people contributed to this project and, knowing it or not, quietly inched the bar higher. To you, David Vaughan, Annie Proulx, Dan Flores, Geri Stayman — and especially Dylan and Matt Stayman—who remind me always that books matter, I am deeply indebted. To Steve Sagin, who on my first day in Saratoga pointed me in the direction of the Red Desert, thanks for the tip. Thanks too to my traveling companions there over the years—Brett Lequercia, Judi Morris, Toby Jurovics, and Rod Laird; and the intrepid Deryl James, Dave Quitter, and Gerald James, who with Annie, Bob Cook, Charles Ferguson, and Dudley Gardner (and Dudley’s crisp memory of lost maps) made high adventure where there might only have been discovery. To Alan Mitchell, John Boyer, Debbie Rusk, Jamie Newman, Brad Carey, Terry and Jimmy Hinkle, and Larry Hicks, I am grateful for access to places and stories about places I would never otherwise know. For six years, I have depended on others’ sharp eyes to keep these pictures on track—people acting as jurors and critics, or as old friends. Fred Baldwin and Wendy Watriss, Jean Caslin, Peter Goin, Susan Moldenhauer, Patrick Nagatani, Kira Pollack, Jim

Stone, Mary Virginia Swanson, Elizabeth Turk, Tim Wride — thanks for the candor. To Julianne Kost at Adobe Systems and Patrick Carr of Carr Imaging, thanks for taking the mystery out of Photoshop without killing the magic. To Oscar Simpson, thanks for helping me get the captions right. To the staff at the El Rio diner in Baggs, thank you for being there. Before the work was completed, a few museums and galleries lent momentum to the Red Desert project by exhibiting and purchasing early versions of the portfolio. To the University of Wyoming Art Museum in Laramie, the International Cultural Center at Texas Tech University in Lubbock, the Witte Art Museum in San Antonio, the Palm Springs Art Museum, and FotoFest in Houston—I remain grateful. For decades of help in honing my heart-eye coordination, my early mentors must take credit: LaVerne Ornelas, Harold Huber, Jim Dow, John McWilliams, Emmet Gowin, and Linda Connor, thank you. And without my friend and pilot, David Worthington, the big picture would have been impossible to imagine. Martin Stupich

3

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 4

Ferris Mountains and Muddy Creek drainage, northeastern Red Desert, Carbon County, Wyoming, 2002

4

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 6

Honeycomb Buttes, northwestern Red Desert, Sweetwater County, Wyoming, 2002

6

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 5

Virga at sunset near Saratoga, eastern edge of the Red Desert, Carbon County, Wyoming, 2002

5

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 7

Echo Overlook, the confluence of the Yampa and the Green, southwestern corner of the Red Desert

7

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 8

Skull Rim at sunset, Adobe Town, Sweetwater County, Wyoming, 2004

8

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 9

Washakie Formation hoodoos eroding, Carbon County, Wyoming, 2003

9

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 10

Sandstone columns below Skull Rim, Adobe Town, Sweetwater County, Wyoming, 2004

10

Proulx_01_pp.001-039:Layout 1

8/21/08

3:22 PM

Page 11

Sandstone lag litters shaly badlands, dawn, the Haystacks, Sweetwater County, Wyoming, 2006

11

Proulx_01_pp.001-039:Layout 1

8/21/08

12:44 PM

Page 12

Weathered potholes, Little Firehole Canyon east of Flaming Gorge, Sweetwater County, Wyoming, 2004

12

Proulx_01_pp.001-039:Layout 1

8/21/08

12:45 PM

Page 13

April snow squall at Sage Creek Gap east of Flaming Gorge, Sweetwater County, Wyoming, 2004

13

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 14

Grove of ancient junipers east of Sage Creek basin, Ashley National Forest, Sweetwater County, Wyoming, 2004

14

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 15

Feral horses near Bitter Creek townsite, Sweetwater County, Wyoming, 2003

15

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 16

Killpecker barchan sand dunes migrating over rim of Great Divide basin between Steamboat Mountain and Black Rock Butte, Sweetwater County, Wyoming, 2002

16

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 17

Boar’s Tusk volcanic plug, Killpecker dune field and automobile dump, Sweetwater County, Wyoming, 2007

17

Proulx_01_pp.001-039:Layout 1

8/6/08

5:26 PM

Page 18

One-meter-diameter nest of ferruginous hawk on isolated bedding remnant, Skull Rim, Adobe Town, Sweetwater County, Wyoming, 2004

18

Proulx_01_pp.001-039:Layout 1

8/21/08

3:52 PM

Page 19

Hind legs, equine carcass, Carbon/Sweetwater county line, Wyoming, 2002

19

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 20

Grasshopper specimens from the Red Desert, University of Wyoming entomology collection, 2006

20

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 21

Cobble-pebble lag concentrated in a dry rill, Adobe Town, Sweetwater County, Wyoming, 2004

21

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 22

Battle Mountain between the Savery and Little Snake valleys, southeast Red Desert, Carbon County, Wyoming, 2002

22

Proulx_01_pp.001-039:Layout 1

8/21/08

3:23 PM

Page 23

Timber-littered alkali shoreline, Soda Lake south of Bairoil, Carbon County, Wyoming, 2006

23

Proulx_01_pp.001-039:Layout 1

8/6/08

Sandstone shaped to accept human hands, White Mountain area, Sweetwater County, Wyoming, 2003 24

5:27 PM

Page 24

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 25

Petroglyph in sandstone, Powder Springs Wash, Sweetwater County, Wyoming, 2006

25

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 26

Cairn near Brown’s Hill, Carbon County, Wyoming, 2005

26

Proulx_01_pp.001-039:Layout 1

8/21/08

3:25 PM

Page 27

Cairn, one of a series, northeast of Dad, Carbon County, Wyoming, 2005

27

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 28

Inscriptions made by nineteenth-century emigrants atop Independence Rock at the point where the California-OregonMormon Trail nears the Red Desert, Natrona County, Wyoming, 2006

28

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 29

Footing at coal mine ruins, ghost town of Dines, Sweetwater County, Wyoming, 2004

29

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 30

Collapsing homestead and hawk’s nest on the Ferris-Bairoil Road, Carbon County, Wyoming, 2006

30

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 31

Remains of the town of Dad, Carbon County, Wyoming, 2003

31

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 32

Skyline of Wamsutter, epicenter of twenty-first-century gas boom, Carbon County, Wyoming, 2005

32

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 33

Thrift shop and liquor store, town center, Baggs, Carbon County, Wyoming, 2005

33

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 34

Downtown Rawlins with old Union Pacific Railroad station in background, Carbon County, Wyoming

34

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 35

Remains of metal shed, Pettigrew homestead north of Rawlins, Carbon County, Wyoming, 2006

35

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 36

Motel Yes No, Jeffrey City, former uranium boomtown known as Home on the Range, Fremont County, Wyoming, 2006

36

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 37

Double cross at grave of Jose Gonsales, possible cholera victim, Mexican Flat, Carbon County, Wyoming, 2005

37

Proulx_01_pp.001-039:Layout 1

8/6/08

5:27 PM

Page 38

Derelict section of old Lincoln Highway near Patrick Draw, Sweetwater County, Wyoming, 2003

38

Proulx_01_pp.001-039:Layout 1

8/22/08

1:04 PM

Page 39

Railroad tracks overgrown with rabbitbrush and sage, Carbon County, Wyoming

39

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 40

Pickup and horse trailer on Cherokee Rim, Washakie Basin, Sweetwater County, Wyoming, 2006

40

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 41

Chain Lakes Flat with trailer, Sweetwater County, Wyoming, 2006

41

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 42

Train derailment, Wamsutter, Sweetwater County, Wyoming, 2005

42

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 43

Overturned and shot-up Pontiac near Soda Lake Flat, Carbon County, Wyoming, 2003

43

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 44

Monolith, lichen, and Eightmile Lake from Atlantic Rim, Carbon County, Wyoming, 2004

44

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 45

Star Wars in roadside trash north of Rock Springs, Sweetwater County, Wyoming, 2004

45

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 46

Playa crust with eyeglasses, Red Creek drainage, Carbon County, Wyoming, 2003

46

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 47

Iron pipe merging with alkali north of Rawlins, Sweetwater County, Wyoming, 2003

47

Proulx_02_pp.040-072:Layout 1

8/21/08

3:27 PM

Page 48

Designated Colorado State Recreation Site for off-road recreational vehicles, Moffat County, Colorado, 2006

48

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 49

Gas pipeline rights-of-way intersecting near Bitter Creek, Sweetwater County, Wyoming, 2002

49

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 50

Maze of access roads in gas fields near confluence of Sand Creek and the Little Snake, Carbon County, Wyoming, 2004

50

Proulx_02_pp.040-072:Layout 1

8/22/08

1:01 PM

Page 51

Pipeline trench right-of-way disturbance east of Adobe Town, Sweetwater County, Wyoming, 2002

51

Proulx_02_pp.040-072:Layout 1

8/6/08

5:33 PM

Page 52

Open-pit uranium mine at Sheep Mountain south of Jeffrey City, Fremont County, Wyoming, 2005

52

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 53

Irrigation reservoir rising behind newly completed earthen dam on the High Savery, Carbon County, Wyoming, 2004

53

Proulx_02_pp.040-072:Layout 1

8/6/08

Mining operation near Flaming Gorge, Daggett-Uinta county line, Utah, 2004 54

5:34 PM

Page 54

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 55

Open-pit coal mine and tailings near Jim Bridger Power Plant, Sweetwater County, Wyoming, 2002

55

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 56

Cooling ponds at the Jim Bridger coal-fired power plant, Sweetwater County, Wyoming, 2002

56

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 57

Plumes rising at Jim Bridger coal-fired power plant near the geographic center of the Red Desert, Sweetwater County, Wyoming, 2002

57

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 58

Seismic monitoring equipment used to identify prospective gas well drilling sites east of Adobe Town, Sweetwater County, Wyoming, 2006

58

Proulx_02_pp.040-072:Layout 1

8/22/08

1:02 PM

Page 59

Newly excavated pit for gas well south of Wamsutter, Sweetwater County, Wyoming, 2005

59

Proulx_02_pp.040-072:Layout 1

8/21/08

3:28 PM

Page 60

Blowout preventer below drilling platform, gas rig north of Wamsutter, Sweetwater County, Wyoming, 2005

60

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 61

Halliburton truck passing, north of Baggs, Carbon County, Wyoming, 2005

61

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 62

Lined pit at drilling site, fenced and flagged, south of Crook’s Gap, Fremont County, Wyoming, 2006

62

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 63

Drilling site aftermath east of Chain Lakes Flat, Sweetwater County, Wyoming, 2006

63

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 64

Steel “flaring tower” guyed in place, former gas pad west of Barrel Springs Draw, Sweetwater County, Wyoming, 2005

64

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 65

Cross-country gas pipeline construction in progress near Grenville Dome, Carbon County, Wyoming

65

Proulx_02_pp.040-072:Layout 1

8/22/08

1:03 PM

Page 66

Surface-laid pressurized gas pipeline, Vermillion Creek drainage, Sweetwater County, Wyoming, 2006 66

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 67

Newly laid forty-two-inch gas pipeline near Walcott Junction, Carbon County, Wyoming, 2006

67

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 68

Forty-two-inch gas pipeline section at staging area near Sinclair, Carbon County, Wyoming, 2006

68

Proulx_02_pp.040-072:Layout 1

8/21/08

4:47 PM

Page 69

Shot-up oil drum between Wamsutter and Crooks Gap, Sweetwater County, Wyoming, 2003

69

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 70

A twelve-meter new mud pot in the Washakie basin east of Dad, Carbon County, Wyoming, 2007

70

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 71

Maze of drilling pad roads west of Rawlins, Carbon County, Wyoming, 2004

71

Proulx_02_pp.040-072:Layout 1

8/6/08

5:34 PM

Page 72

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 73

PART II text

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 74

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 75

acknowledgments

Very many historians and researchers helped with information and connections with other people who knew something pertaining to the work in hand. All thanks to photographer Marty Stupich, whose idea of a book on the Red Desert started this project. Dr. Fred Lindzey, retired professor in Zoology and Physiology at the University of Wyoming, was the keystone connector to the naturalists at the Wyoming Natural Diversity Database at the University of Wyoming. Deputy Mary M. Hopkins at the Wyoming State Historic Preservation Office, and Steven Sutter, cultural resource specialist in the same organization, rooted out yellowed archeological reports that shed light on some of the mysteries of the Red Desert. The entire staff of the University of Wyoming’s American Heritage Center helped in many ways over the years in ferreting out rare manuscripts and ephemeral references related to Wyoming’s past. Leslie Shores, photo archivist, gave substance and place to the shadowed histories of early Wyoming ranch life. Assistant reference archivist Anne Guzzo provided a copy of the film Fight of the Wild Stallions, showing Red Desert horse-catching techniques in the 1940s. Trina Purcell of the Denver Public Library’s Western History and Genealogy department solved the riddle of the mysterious “CWA” cited often as a source in John Rolfe Burroughs’s Where the Old West Stayed Young. This turned out to refer to the Civil Works Administration, whose valuable 1930s interviews with residents of Colorado counties are housed in the Colorado Historical Society archives. Dave Quitter of Saratoga spent many hours working to find references to Quien Hornet, the name of a mountain in the Red Desert that appeared on Howard Stansbury’s map. In the end,

weeks of work resulted in a single footnote. He also helped with fine-tuning the geological photograph captions. Dan Davidson, the director of the Museum of Northwest Colorado in Craig, pointed us to L. H. “Doc” Chivington’s useful manuscript, “Last Guard,” perhaps the best account of an ordinary cowpoke’s job at the turn of the last century in southwest Wyoming–northwest Colorado. Jan Gerber, the assistant director of this museum, provided photographs and information on bygone personalities of the Little Snake River valley, including W. W. “Wiff” Wilson, who bridged the era between cattle raising and oil extraction in Wyoming. Cindy L. Brown, the reference archivist at the Wyoming Division of Cultural Resources, was helpful in tracking down references to the name change of “Washakie” to “Wamsutter.” Sandra Lowry, the librarian at the Fort Laramie National Historic Site, helped with day orders and documents showing that “Fort” LaClede was not, in fact, part of the military fort system in Wyoming, but one of the Overland Stage stations fortified by Holladay. Marva Felchlin, the director of the Autry Library at the Autry National Center in Los Angeles, was very helpful with audio disks of lectures on American violence, especially in the west, and materials related to Wyoming outlaws and the diary of Second Lieutenant William Abbott, who served at Fort Fred Steele in 1872 –1874. Lindsay Ricketts made the first pass at copyediting and formatting a gnarly tangle of pages. Our rigorous editors at the University of Texas Press earned our profound gratitude for their 75

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 76

punctilious work. Thank you, Megan Giller and freelancer Rosemary Wetherold. Throughout the study period, the Western Wyoming Community College Oral History Project interviews with many people now gone helped us gain an intimate view of those who lived in and around the Red Desert. The useful bulletins and updates from concerned organizations included the Wyoming Outdoor Council, the Biodiversity Conservation Alliance, and Friends of the Red Desert. Bob Cook, outdoor liaison, map interpreter, and gifted selector of the best camping spots was indispensable on trips into the desert. Dudley Gardner was a spark plug; his terrible energy, unfailing optimism, and gritty determination, his knowledge of good coffee sources in southwest Wyoming and northwest Colorado, and his willingness to drive long, bad roads and sleep on the truck seat made him the ideal historian-archeologist for the project. Pam Murdock and Mary Wilson of the blm Rawlins Field Office guided us through the historically important JO ranch and were very helpful in supplying historical reports on the JO and Jawbone ranches. Thanks also to Oscar Simpson of the New Mexico Oil Conservation Division and the New Mexico Wildlife Federation for explanations of coal-bed methane gas extraction machinery and procedure. Russel Tanner commented that rock art scholars “who deserve mention from the beginning for their contributions to the greater endeavor of scientific inventory, context development and description of various Native American rock art sites include David T. Vlcek, James Keyser, George Poetschat, Julie Francis, Alice Tratebus, Larry Loendorf, Linda Olson, Dudley Gardner, Bill Current, Tom Larson, Ron Dorn, Sam Drucker, and Joseph Bozovich Jr., and his esteemed father, the late Joe Bozovich Sr.” Russel Tanner also wishes to thank Clifford Duncan, Starr Weed, Floyd Osborn, Haman Wise, Delphine Clair, Burton Hutchinson, Henry Antelope, Bobby Joe Goggles, Sherry Blackburn, and Diana Mitchell, as well as the late Shorty Ferris, John Tarness, and Joe Pinnecouse. 76

George P. Jones thanks Fred Lindzey and Marty Stupich. Jeffrey Lockwood wrote that “his contribution to this book would not have been possible without the assistance of Scott Schell, Gary Beauvais, and George Jones.” Charles Ferguson thanks Andrew Coen and Ron Surdam for discussions regarding limnology. Gerald Smith and Jon Spencer shared valuable unpublished information regarding fish evolution and the Colorado River. Richard W. Jones was very helpful with reference materials, as was the entire staff of the Brinkerhoff Geology Library at the University of Wyoming. Robert L. Cook was an invaluable field assistant. Annie Proulx, Steve Cather, and Michael Mahan reviewed earlier versions of the manuscript and improved it immensely. Laura and Walter Fertig would like to thank Gary Beauvais, George Jones, Ron Hartman, and Bonnie Heidel for sharing thoughts and input on the ecology and botany of the Red Desert. Andrea Orabona offers “sincere thanks to naturalists Mac Blewer, Marian Doane, and John Mionczynski for sharing their expertise and experiences with me while conducting additional research” for her chapter on birds of the Red Desert. Dudley Gardner thanks Annie Proulx for “listening and editing” and writes that “Jana Pastor, David Johnson, Danny Walker, Mary Lou Larson, Michael Metcalf, Russel Tanner, Emma Gardner, Richard Etulain, Ken Fitschen, Barbara Smith, Murl Dirksen, Martin Lammers, Will Gardner, and Byron Loosle are owed my gratitude for commenting on the technical aspects of the sections on environment and prehistory.” Craig Thompson states, “I regret the passive neglect suffered by my children, my wife, Jocelyn, and my students during my frenetic research and writing periods. I acknowledge the invaluable help of Dr. Ron Surdam, master detective and premier chronicler of Lake Gosiute, and Charles Matheson Love, the most observant fieldman with whom I have ever worked.”

Acknowledgments

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 77

introduction

This book is not intended as another plea to save the greater Red Desert. Many tries for conservation by people who love the place have come and gone over the decades, defeated by the prevailing attitude of “show me the money,” by the congressional cold shoulder, by lack of knowledge of what is in that high desert, by the complex mixture of politics and culture, and by the momentum of our times, inexorably propelled by shifting global histories, which, like massive continental plates, have thrust us into the present. This book tries to sort out what there is about the Red Desert that makes it valuable, scientifically and historically interesting. We hoped to dispel some of the myths that have grown up around the place. There is nothing in the Red Desert but sagebrush, and who needs it? Now that forage is depleted, the Red Desert is empty wasteland. Coal, oil, gas, gravel, and trona are its only assets. There were never any ranches in the Red Desert. Wildlife is mostly feral horses that eat grass that could otherwise nourish cattle and sheep. There is no water. The Red Desert had nothing to do with the formation of the Grand Canyon in Arizona. The project began several years ago when photographer Martin Stupich asked me to write an introduction to a collection of his photographic work in the Red Desert. I agreed, thinking it would be a simple matter to go to the University of Wyoming’s library, gather up the books on the region, and write a general overview of the place. I was stunned to discover there was not one single book on the Red Desert in Coe Library. Nor did a search of

the American Heritage Center’s archives turn up much beyond an old photograph of a locomotive stalled by the infamous 1949 blizzard, and a nineteenth-century paper on forage plants by the esteemed botanist Aven Nelson. What we did not know about this huge area of the state began to swell up into a thundercloud of general ignorance. Martin Stupich’s photographs were not meant to illustrate the text, but are his stand-alone record of the desert over a period of years. The text grew up around what we didn’t know rather than the photographs. As we worked on the book, we learned that Red Desert fieldwork and study could absorb many lifetimes. Our book is only a start. Almost everyone who worked on the project came away excited by what they found, dazzled by the possibilities for learning more. Hydrologist Craig Thompson mentioned that Bitter Creek, draining one of the largest basins in North America, could enrich our knowledge of water in high desert.1 There are flowing wells out there that serve as oases for wildlife, ice lenses in the sand dunes that melt into pools of water. None of these are well studied. Prospector Bob Cook said that every trip revealed something new to him—barrel hoops, purple glass, selenium indicator plants, and “a zillion roads; it’s a spider web out there.”2 Entomologist Jeff Lockwood was amazed by the riches of insect life, which he said seemed “absolutely staggering,” adding that the diversity “has got to be phenomenal . . . the place is not homogeneous.”3 We learned that the Red is composed of rich pocket habitats catering to a wide variety of specialized vertebrates, invertebrates and plants, pockets that are scarcely known and certainly not mapped. Lockwood described these habitat pockets as 77

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 78

Cautions on Venturing into the Desert annie proulx The Red Desert is one of those places that easily humble and even humiliate those who think they know it. No one person knows this place. Many people know special small corners, feel at home on the flanks of certain mountains or inside the particular maze of rock outcrops or up on Green Mountain or the Haystacks looking south into the desert, but weather, rockslides, reworked roads, and sidetracks can disorient even old hands. Readers who are curious to see the Red Desert for themselves should be careful. If possible, do not go alone, and if you have never been there before, go with someone who knows the way. If you are alone and injure yourself, you may find yourself in a bad situation. Some roads are thickly strewn with flinty rocks that can shred flimsy tires with ease. Flat

tires are a common mishap in the Red. Ornithologist Andrea Orabona recalled what happened one year when, just after she had had knee surgery, she was doing her bird survey in the Red Desert: “On that trip, I was in my Game and Fish truck and got two flat tires on the way out of the Red Desert, but had only one spare. I always take my trekking poles, but also take my mountain bike—just in case—if there’s room. Since I was camping out of the back of my truck, there wasn’t room for the bike that time, so my only option was to walk out. I knew where the Sweetwater was.”1 Luck was with her. She met some lost tourists who had been looking for a grave site. They took her back to Lander, and she answered their many questions about wildlife and the desert. “It was a mutual rescue,” she said. Geologist Charles Ferguson and prospector Bob Cook had a vehicle breakdown and faced a walk of a few miles to a larger gravel road that

archipelagos of life in the sea that is the Red Desert. The scale of seeing is very different as well, said Lockwood— ”so much of what happens in the Red Desert happens within ten feet and beyond a thousand feet.”4 One constantly shifts perspective from what is close at hand to the far horizon line, creating a kind of psychological double-think, constantly forcing one to consider the particular in relation to the whole. Dotted over the desert are old ranch and even ghost town sites rich in dumps—tin cans, machinery, appliances, cars, all the bits and pieces of old-time Wyoming life. Russ Tanner mentioned that the mining ghost town of Sublet has a dump of cans that covers three hundred acres. Can historians, archeologists, and their students learn something from the detritus of the western yesteryear? Botanists Walter and Laura Fertig fell in love doing fieldwork 78

might have traffic on it. There was no traffic, and they waited and waited for many hours before someone—a Sweetwater County road grader— came along. A rugged, four-wheel-drive vehicle with high clearance is a necessity on the old tracks, and heavy-duty tires an asset. Most of the older roads are dry-weather roads. It is possible to get caught in a passing summer shower and feel the road turn to lard beneath your wheels. Desert rats keep a copy of War and Peace in their vehicles to read while the road dries out. What looks like solid ground off-road can, if you drive onto it, reveal its true character as a bottomless bog. The new, wider gravel roads built by and for the cbm (coal-bed methane) extraction companies generally are one-ways to rigs and compression stations. They may have signs, they may not. The multiplicity of new roads and reworked roads makes it easy to get enmeshed in a cat’s cradle of connecting

in the Red Desert a decade ago. Walter Fertig is a specialist in Wyoming’s rare plants, and he wrote, “It was such a tremendous personal discovery to find so many unusual and interesting species in places that so many people thought were without any redeeming value—just wastelands only good for drilling oil and gas, feeding cows, or driving across really fast.”5 He continued: “And as a conservation-minded naturalist, I can see that the Wyoming basin country really is the area that we are most responsible for conserving into the future. Of course a lot of my friends and colleagues think I’m a little crazy for holding this view, and it has been a frustrating, uphill effort for years trying to convince other conservation-minded individuals and groups that lands like the Red Desert need to be protected as much as, if not more than, the Yellowstones.”6

Introduction

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 79

tracks that seem to go nowhere. In other places the old tracks have been obliterated by pipeline and new road construction, and finding your way to a specific area may be possible only with a guide, close study of several maps, and/or GPS guidance. And even recent maps have little relation to the reality of the gasextraction present. Landmarks are as important today as they were when Stansbury crossed the desert. If you are hiking in canyon or malpais areas, thirst and heat can drop you to your knees. It is easy to get lost in an immense place where rock outcrops obstruct the sightlines. Check ahead as you walk, and check your back trail often for landmarks. Cell phone reception is a joke. In summer it is hot and dry unless you know where the water is. There are no cafés or refreshment stands in the desert. It is smart to bring extra water, gas, tools, maps, gps units, and a spare tire or even two.

Certain organizations—the Biodiversity Alliance, Friends of the Red Desert, and the Wyoming Outdoor Council—sometimes offer guided trips into the Red Desert, and this may be the best way to first experience the big and difficult place. Trial-and-error exploration can take many years. Do not remove arrowheads, scrapers, or other artifacts. The Red Desert has many archeological sites. Often one stumbles on places with a good view where long ago someone sat making arrow points and stone tools. Notify the state archeologist of locations and/or gps coordinates if you discover rock art, areas of stone tools, and such. One year we noticed a certain hunting camp in the Washakie basin. Over the years hunters had found and collected many stone points and tools. They removed them from the sites of origin and brought them to the hunting camp and arranged them on a flat rock apparently as

As entomologist Jeff Lockwood pointed out, millions of people pass through the Red Desert every year without seeing it. Interstate 80 bisects the desert, and trucks and cars rush past seemingly endless miles of sagebrush, dismissing what they see as monotonous and useless. From that highway the Red Desert does not seem interesting. This is an illusion, for this place is rich in fossils, a vast stone book of pictographs and artifacts of Native American tribes. The early explorers of the west came through it. It was part of the great westward emigration. The Union Pacific Railroad cut through its center, separating the bison into north and south herds. Great sheep trails from California to the east and New Mexico to the north traversed the Red Desert. Bison, desert elk, pronghorn, and bighorn sheep lived in it. Fish swam in its waters. Outlaws hid in its folds and badlands. And a num-

some sort of hunting shrine. Archeologists, now unable to make connections of those tools with their original sites, photographed that collection of artifacts in 2004 and periodically check them. At least the hunters did not carry them away. To bring a gun or not is a personal preference. On our first trip into the desert, we came upon a pronghorn lying on the bank of a new energy road. It had been hit by a speeding vehicle and its pelvis was broken. Cell phones were useless, and had not one of the people in our group had a gun, the animal would have died a slow, lingering death. Watch the sky for weather changes.

1. E-mail message to the author, July 23, 2007.

ber of hardy people, abetted by the Homestead Act of 1862, impressed by the big open space, and hungry for land—many of them ex – railroad workers and coal miners — tried to establish and maintain ranches in remote and difficult corners. Some succeeded, and all contributed to the state’s reputation for hardiness, cheerful toughness, and ability to withstand privation. It is our hope that this book will encourage naturalists, historians, graduate students, and Wyoming residents to venture into the Red Desert and discover for themselves the microhabitats, curiosities, and beauty of what remains in this little-known place, that they will observe for themselves the new roads and attendant dust storms, notice the biomass of halogeton, Russian thistle, cheatgrass, and other invasive weeds along those roads that come with soil disturbance. It is easy to blame all the changes in the

Introduction

79

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 80

Red Desert on energy extraction work, but that is the narrow view. There are countless Red Deserts in this world. Jack States touched the larger problem when he said, “Undeniably much of the pristine Red Desert ecosystem is imperiled not only by resource hungry corporations fueled by a resource hungry populace (that includes sanctimonious environmentalists), but also by inexorable global warming and extinction of species. To me the issues we face in the Red Desert are not that different from any other aspect of global environmental crisis spawned by a burgeoning human population.”7 There are important subjects that neither time nor space allowed us to include: a review of fossil finds from the Fossil-UintaGosiute lakes; fish, extinct and living; ice lenses in the Killpecker Dunes; fire in the Red Desert; soil studies; wind, weather, and the severe drought that continues to grip the entire west; the errorriddled survey of the southern state line, which caused land ownership problems for years; the history of the dryland farm colony on Brown’s Hill in the 1920s; hunters and poachers; the people of the North Country and the Sweetwater Valley; the labor history of the coal mining towns; the extraordinary scenery of the badlands. Nor was there room for the stories of characters and tough old desert rats like the hermit who trimmed his hair with a lighted newspaper; horse catcher Tex Love, found sitting dead against a rock; the fellow who found the frozen body of a dead sheepherder, loaded him onto the roof of his vehicle, and casually drove around the nearest town doing errands. The women freighters, the practice of boarding ranch children in town for schooling, and the medical exigencies in sickness and accident all deserve examination. Someone who spent time in the Red Desert years ago recently remarked to me that he could not bear to go back and see it in a spoiled condition. But we have to go back. We cannot turn away from the place. Jack States, who wrote on the biological soil crusts for this book, was born and raised in Wyoming. He related: “In my younger days I was strongly influenced by my grandfather who gauged the land by how many sheep it would support while at the same time communicating to me his deep concerns regarding va80

garies of environment & weather and the cumulative negative effects of overgrazing. It was out of necessity that he was a steward of natural resources, but it was out of respect and love for creation that he strived for balance, cultural with natural, conservation with exploitation. Because I value the beauty and solitude of the Red Desert landscapes, I admit to being intensely possessive, even angry at the fresh tire track, the atv, and the army of seismic trucks. But I also believe that with good science, we can successfully manage its natural resources for sustainability.”8 For Gary Beauvais the great characteristic of the Red Desert is “sheer open SPACE.” He wrote: “Boil down that piece of country to its core, and outright vastness is left. . . . Space is responsible for many of the Red Desert wildlife values. The Red Desert is a stronghold for greater sage-grouse because they can lek in one place, nest thirty miles away, fly another twenty miles to primary winter range, then move another twenty to survival range if the weather gets really bad. It’s a stronghold for low-country elk because they can use sheer distance to avoid disturbances. I’ve spooked herds around Steamboat Mountain and watched them run, steadily, for at least a couple of miles before they fade out of sight behind a distant ridge. A golden eagle perched on a Red Desert butte or rim surveys more country than any eagle ever did sitting on a power pole in the Great Plains. And every road, well field, and pit mine makes it smaller. Whereas before there were no boundary features, now we are carving all sorts of lines into the Red Desert. Some are visual only . . . like 2-tracks and abandoned railroads . . . but others are functional, actual boundaries, like I80 and Jonah Field and coal-bed methane fields with their attendant roads and man camps.”9 The many-faceted character of this high, cold desert haunts those who have spent time in its irreplaceable silence and space. Mac Blewer, living in the Washington, D.C., corridor (“Land of Perpetual Self Importance”), knows that. He wrote: “The desert comes to me now in my sleep. The other night I dreamed of antelope bounding through the sage and rabbit-brush. A murky shape that looked like Steamboat Mountain— covered in fog—

Introduction

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 81

was in the background. I remembered the desert looking like that after a blizzard in late June several years ago. I had been hiking . . . in the Sands and had been surprised by the brief wintry storm that had caught me unawares. After 24 hours of snow, hail, wind and lightning, the sun came out, the snow melted, and a thick mist rose up. Walking through the aspens and limber pine, I could see new flowers sprouting and butterflies sunning themselves on snow-banks. After a time the fog lifted and I walked in peace and awe on the top of this desert island, wishing that my sojourn would never end.”10 The Red Desert needs to be examined and studied. Perhaps the most we can hope to save are small pieces, little corners of what was once the largest area of unfenced land left in the United States, forced to ignore the old truth that the whole is greater than the sum of its parts. Something ineffable is already lost. But Andrea Orabona remarks that there could be great value and utility in preserving the Red Desert as a national conservation area.11 It would allow us to study a place about which we know very little, and although there are a dozen of these areas in the United States, there is no national conservation area in Wyoming. They are designated by Congress to “conserve, protect, enhance, and manage public land areas for the benefit and enjoyment of present and future generations.” Such areas “feature exceptional natural, recreational, cultural, wildlife, aquatic, archeological, paleontological, historical, educational or scientific resources.”12 The Red Desert meets every one of these criteria. In November 2007, Adobe Town gained partial protection from noncoal surface, uranium, and oil shale mining in a 5–1 vote by the Wyoming Environmental Quality Council. It is not protected from oil and gas extraction. It is sad to end on a bitter note, but as we were finishing this book, we learned that vandals had slashed archeologist Dudley Gardner’s tent with knives and damaged his dig site near a new and oncoming pipeline installation.

Notes 1. Pers. comm., June 9, 2007. 2. Pers. comm., June 9, 2007. 3. Pers. comm., June 9, 2007. 4. Pers. comm., June 7, 1007. 5. E-mail message to the author, May 31, 2007. 6. Ibid. 7. E-mail message to the author, June 8, 2007. 8. Ibid. 9. E-mail message to the author, June 7, 2007. 10. E-mail message to the author, June 21, 2007. 11. Pers. comm., June 7, 2007. 12. National Landscape Conservation System, http://www.blm.gov/ nlcs/conservation/index.html.

Annie Proulx, Saratoga, Wyoming Introduction

81

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 82

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 83

natural history

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 84

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 85

1. Geology of the Red Desert Charles Ferguson

Prelude Steamboat Mountain is for many the most important landmark in the Red Desert. The lava-capped mesa 50 kilometers north of Rock Springs towers over the desert and forms the northern rampart of a prominent wind gap, 400 meters deep, through which spill the Killpecker Dunes. Why is Steamboat so high, and what’s with an active dune field more than 100 miles long, but only 1 to 3 miles wide? “Be sure to visit Steamboat Mountain, and check out the dunes” is what geo-anthropologist Charlie Love advised me when I started this project. Killpecker Creek was a problem too; the south-flowing ephemeral wash that empties into Bitter Creek at Rock Springs occupies a valley fit for a king. It was stormy, late in the field season, and already dark when I finally got to Steamboat and set up camp. I walked to the rim to watch the moon rise over Black Rock Flat. Standing on the rim, leaning into the wind with shreds of cloud tearing by at eye level, is like standing on the bow of a huge ship plowing through an invisible sea, dunes sprawling out across a terrifyingly visible abyssal plain that stretches out forever in all directions. Later I grilled a steak in a grove with blue moon glare casting sharp sage and twisted limber pine shadows. While I ate, a series of squalls streamed around the mountain, enveloping the camp in Kuro-

sawan fog banks, the fire, virtually invisible in the moonlight, springing to light each time. A Hard Place to Get To The Red Desert as a geologic entity is here considered to include the 10-million-acre expanse of southwestern Wyoming that encompasses the Green River, Great Divide, and Washakie basins.1 There are very few places on earth where rocks are older, but none have such a history. This is because the Red Desert is also exceptionally young. Its Cenozoic record is like no other on earth (see Table 1.1),2 and even though none were there to bear witness, its landscape has changed dramatically since the dawn of man. When the Red Desert was first being explored at the start of the nineteenth century, modern geology, the science of deep time based on James Hutton’s radical new theory that the earth was very old, was finally maturing.3 So it was that exploration and science converged, making the Red Desert one of the few places on earth whose geologic wonders were first viewed by modern geologic eyes. The white man’s discovery of fossils, coal, and oil, all known to Indians, made the Red Desert a focal point for the testing of a young science’s crucial theories ranging from evolution to the origin of igneous rocks. 85

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 86

Table 1.1. Geologic time—the eons, eras, and periods of earth history and the epochs of the Cenozoic (in Ma) Eon Era Period Epoch Phanerozoic 0 –540 Cenozoic 0 – 65 Quaternary 0 –1.8 Holocene 0 – 0.01 Pleistocene 0.01–1.8 Neogene 1.8 –24 Pliocene 1.8 –5.3 Miocene 5.3 –24 Paleogene 24 – 65 Oligocene 24 – 34 Eocene 34 –55 Paleocene 55 – 65 Mesozoic 65 –248 Cretaceous 65 –144 Jurassic 144 –206 Triassic 206 –248 Paleozoic 248 –540 Permian 248 –290 Pennsylvanian 290 – 323 Mississippian 323 – 354 Devonian 354 – 417 Silurian 417– 443 Ordovician 443 – 490 Cambrian 490 –540 Proterozoic 540 –2,500 Neo 540 –900 Meso 900 –1,600 Paleo 1,600 –2,500 Archean 2,500 – 4,000 Late 2,500 – 3,000 Middle 3,000 – 3,400 Early 3,400 – 4,000 Age of Earth 4,550

86

Note: Including the slopes of its flanking mountain ranges, outcrops of every age of the earth can be found in the Red Desert, making it possibly the only place on earth where this is true. Ages in this table, based on radiometric dating, are in millions of years (Ma). Geologists denote deep time in three bunches of triple zeros: thousands = ka, millions = Ma, and billions = Ga (kilo-, mega-, and giga-annum, respectively). Using this notation, the earth is 4.5 Ga, 4,500 Ma, or 4,500,000 ka. The average expected life span of a human is 0.075 ka. The oldest rock in the world, using the same notation, is 3,960,000 ka (Bowring et al. 1989). Gunter Faure (1977) provides a brief but succinct history of radiometric geochronology. The 14C system, based on an isotope with a very short half-life, can be used only to detect recent time—say, within the past 50,000 years or so. For deep time two main techniques, K-Ar and U-Pb, are used. U-Pb systematics can be measured in minerals that are very resistant to resetting during intense episodes of heating and alteration. The K-Ar system, supplanted in recent years by the more precise 40Ar/39Ar technique, is more sensitive to reheating, a fact geologists take advantage of to chronicle not only the absolute age of minerals but also the cooling history of rocks and ranges.

Charles Ferguson

Page 87

Ne

w

et

Swe South Pass

ter

wa

CM

GM

2624

ol di er

y nd Sa

ee

n

Ha

ms

2070

2090

2100

2080 2105

2070 2150

2075

r er C Bitt

t

Steamboat Mt 2647 2228 Black Rock 2133

ul fa

G D B

2090

1985

Gr

2130 2220

tS

2160

2730

s Lo

2135

G R B g Bi

2530

2250

2067

2078

2172*

Creston Junction

2060*

2070*

1900 2040*

RSU

1850

2050

WB Blue Gap 2010

nd

Cr

Sa

dy Mud

k For

Cr

1900

1850 Quaternary alluvial-lacustrine / eolian

ke

a

Sn

Pleistocene river gravel Neogene volcanics

Litt le

After the Civil War, three major geologic expeditions converged on the Red Desert: Clarence King up the Snake and its tributaries from the west (1867–1871), John Wesley Powell from the south in 1868 and then down the Green in 1869, and Ferdinand V. Hayden (1867–1868) up the Platte from the east.4 Each sought, much like the rivers they had ascended, to wrest control and authority from the others. Naming rights, credits, and vanities were major concerns at the time. The quality of their work is what remains. Then came the serious fossil hunters. Their work galvanized opinion among a select group of scientists who were rigorously testing a new hypothesis of evolution.5 Cenozoic vertebrates, unlike the tiny invertebrates encased in limey shoals, were like us, warm-blooded, furry, familiar, and fascinating. Might there be evidence of human evolution in the bones scattered across the Red Desert? Samuel Knight wrote: “Leidy, Marsh, and Cope, followed by Osborn and Scott, went after the vertebrates. The reverberations of the famous Marsh-Cope feud, which appears to have originated over rivalry for Basin vertebrates, has echoed and re-echoed over the badlands for nearly a century. Thus we see geologic titans tilting with ‘groups’ and ‘species’ in a tournament of rocks and fossils until the Congress, who was, for the most part, paying the bills, rewrote the rules and established the United States Geological Survey.”6

Neogene sedimentary Bishop Conglomerate

en

6:17 PM

Bridger - Ice Point - Crooks Gap

Gre

8/4/08

Killpecker Cr

Proulx_03_pp.073-120:Layout 1

Eocene Wasatch / Green River Paleocene Upper Cretaceous Mesozoic

Yamp a

Fault

Paleozoic Precambrian

0

20

0

Continental Divide

40 20

60 40

80

100 km 60

80

100 miles

Figure 1

A Global Perspective If Hutton set the course of modern geology, Einstein put it on the fast track of the nuclear age. About the same time that isotope geochemists were figuring out how to use nuclear technology to date rocks, oceanographers learned how to map the ocean floor.7 Geologists were confronted with an entirely new mountain range, a midocean ridge rising from the floor of every sea. It soon became apparent that it was not only the youngest, and oldest, but also the world’s longest mountain chain, like the dorsal spine of a Midgardian serpent stretching 50,000 kilometers from a slen-

der Arctic tail to a fat, East Pacific head.8 Youngest because the continuous spine is an undying fissure volcano that adds, on average, 10 centimeters of new crust to the earth every year.9 Oldest because the volcano has been erupting continuously for at least 2 billion years.10 Seafloor spreading is balanced by subduction.11 It has to be. The earth is neither shrinking nor expanding,12 and space must be made. Something has to go back into the mantle. And so it does, but not quietly. The driving force of orogenesis is the push-

Geology of the Red Desert

87

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 88

ing, resisting, shearing, degassing, and sweating that occur at convergent margins.13 The world’s other mountain chain, its convergent orogeny, consists of two strands that meld together in the plate-tectonic mess of fish that is Southeast Asia. The composite Tethyan orogen of the eastern hemisphere is the shorter strand, at 14,000 kilometers, but it is also the highest and widest.14 Its ranges, including the Atlas and the Pyrenees, the Caucasus and the Himalaya, stretch from Casablanca to Singapore and owe their existence to the northward convergence of Africa, India, and Australia with Asia. The longer strand, 45,000 kilometers, circles the Pacific from Erebus to New Zealand and owes its existence to the consumption of oceanic crust around the periphery of the Pacific Rim.15 Its longest uninterrupted stretch, 20,000 kilometers from Tierra del Fuego to Katmai, is the cordillera of the Americas, a complex chain of overlapping and incessantly interacting fold-thrust belts, thick-skinned foreland uplifts, volcanic arcs, transcurrent ranges, and extensional belts. Fold-thrust belts are direct results of continental convergence, that is, head-on-collisions. The closing of the Tethyan Sea, on the other side of the planet, resulted mostly in continent-continent plate collisions. The Pacific Rim, on the other hand, is a convergent margin that pits ocean versus continent.16 The oceanic plate subducts, and the leading edge of the continent crumples. Seaward thickening wedges of continental sedimentary rock get scraped off and thrust up and over the crystalline interior. If any strata resist along the way, they get folded, and then sometimes again just for good measure.17 Fold-thrust belts therefore occur clustered around continental margins. One such mountain belt bounds the Red Desert to the west, and its unusually great distance from the Pacific, its thrust source, is because a younger and exceptionally wide extensional belt, the Great Basin of the western interior, superseded it.180 Farther inland, thickening and shortening of the crust resulted in highly concentrated zones of thrusting. Discrete zones 88

of crustal weaknesses inherited from structures left over after billions of years of orogeny are reactivated and soak up the strain. Deep faults thrust old crystalline rocks up into sharp-margined foreland uplifts.19 The Red Desert’s northern, southern, and eastern bounding ranges are all of this type, and their orientations and shapes are as different as their complex geologic histories. Every age of the earth is represented, and practically every kind of rock known on earth is found there, from commoners like granite and sandstone, to the oddball sodic evaporites of the Green River basin, primordial taconites of the Wind Rivers, and the weird and wonderful wyomingites of the Leucite Hills.20 Volcanic arcs, also referred to as magmatic arcs, are the result of subducted, water-soaked oceanic slabs that literally steam the overlying tectonic plate into submission. The rising fluids, driven out of the oceanic plate as it gets sucked back into the mantle, provide the flux that the already hot, deep crustal rocks need to melt. The buoyant magma rises, some erupting to form volcanoes, but much more solidifying in the middle and upper crust to form monstrous granite batholiths that sit quietly, waiting for the opportunity, often afforded to them during much later and unrelated orogenies, to bob to the surface.21 Zeroing in on a Wide Target Averaging 300 kilometers in width, the American cordillera swells to greater than 800 kilometers in only three places. The lesser two are bulges in central South America (800 kilometers) and northern Canada (1,000 kilometers). The Altiplano-Puna Plateau of the interior Andes, with its monstrous 7-kilometer peaks buried up to their shoulders in their own sediment, are the western hemisphere’s highest desert, and also its highest and largest great divide basin. In stark contrast are the deep, deathlike valleys of the Mackenzie Mountains, flanked by seemingly endless chains of hogback ranges cascading down to the opposing sea.22 The widest place on the world’s longest mountain chain is a bulge that starts in central Mexico and gradually expands north-

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 89

ward to a point where, between San Francisco and Spearfish, South Dakota, it reaches 1,700 kilometers. The bulge corresponds to where the world’s midocean ridge is subducted and overridden by the American cordillera. This is literally the underlyi0ng reason for the anomalous elevation and width of the cordillera that make the Red Desert and its environs so well known. Mountain belts worldwide are well known for their lack of drainage to the sea. Standing high, bordering the arid interior of their host continents, great divide basins (that is, areas that do not drain into the ocean) are to be expected, and so it was believed that a vast inland sea had existed in the interior of North America.23 The sea drained to the west via the Rio San Buenaventura, a mythical river that was supposed to cross the Sierra Nevada of northcentral California and empty into the Pacific at San Francisco Bay. In 1776 the Dominguez-Escalante expedition reached the Green River of northern Utah, proclaimed it to be the Buenaventura, and did not follow it far enough to prove that it was not.24 Because its existence was a key argument in favor of continued investment in western economic ventures, the river persisted on maps for decades after it was known by western explorers to be false. In 1829 Peter Skene Ogden, a Canadian Hudson’s Bay partisan, traversed the east slope of the Sierra Nevada from the Columbia River to the mouth of the Gila River in southern Arizona and proved that no such river could exist.25 Jedediah Smith, the first to cross the Great Basin from east to west in 1826, found no evidence of the river either.26 Joseph Rutherford Walker traversed the Great Basin and the Sierra Nevada in 1833, discovered Yosemite Valley, and also reported that the Buenaventura was a myth.27 It was not until the mid-1840s, after John Charles Fremont also failed to find the elusive river,28 that the mapmakers finally started to erase the Buenaventura. The Wyoming Craton The Red Desert is part of the Wyoming craton, one of a handful of Archean-aged sialic nuclei that form the cores of the world’s

continents.29 They record more than half of earth’s geologic history, yet they make up only a tiny percentage of its crust.30 Most lie tucked away safely within the interiors, insulated from the effects of 2 billion years of continent-building convergent orogeny. The Red Desert is exceptional in that it is the only place in the western and northern hemispheres where Archean crust figures so prominently in mountain building. Archean granites of the Wind River Mountains form the northern boundary of the desert and are the only rocks in this part of the world so old to be thrust so high. In fact, until very recently, Gannett Peak, at 4,207 meters, would have easily qualified as the planet’s highest Archean rock. You have to go deep into the heart of darkest Africa to find a more extreme situation.31 During the late Archean and Paleoproterozoic (3.0 –1.6 Ga), the cratons, which up until then were for the most part wandering freely about the globe, merged together in groups of two or more to form the cores of the continents we recognize today.32 North America was born of the fusion of three—the Wyoming, Slave, and Superior, each forming the point of a monstrous equilateral triangular shield centered on Saskatoon.33 Plate tectonics seems to have really got going, snowballing, so to speak. The bigger the continents got, the more frequently they collided, each time adding new fold-thrust belts and magmatic arcs. During the final 300 million years of the Paleoproterozoic, the world became crowded with crust. The cratons were sutured together by massive, incredibly wide mobile belts, and the Wyoming craton in the process suddenly became an insular terrane surrounded by more than 1,000 kilometers of new crust. What is now the southern Great Plains, most of southern Alberta and northern Montana, and all of Colorado, New Mexico, Utah, and Arizona, was added to the continent at this time. The late Archean and early Proterozoic orogenies of the Wyoming craton are important because their structural grain is largely responsible for the unusual east-west orientation of the Red Desert’s two largest foreland uplifts. The late Archean part of the Wyoming craton corresponds very closely to where the Red Desert

Geology of the Red Desert

89

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 90

is today. It was, in fact, the last part of the craton that was added while it was still a free agent. This happened by an amalgamation of magmatic arcs plastered along what was then the southern margin of the protocraton, a line essentially parallel to the composite Wind River–Granite Mountains foreland uplift.34 The complex fault systems of the northern Red Desert—the Wind River, Continental, Flat Top, Mormon Trail, Crooks Gap, Emigrant Trail, and south Granite Mountains faults — are all inherited from these events. Jadeite, an indicator of high-pressure metamorphism and evidence of Archean subduction, is found there.35 Farther south the slightly younger Cheyenne mobile belt also strikes east-west.36 This strongly sheared southern boundary of the craton is still, after nearly 2 billion years, the principal structural feature that defines the southern boundary of the Red Desert.37 The high-angle faults responsible for the Powder Rim and the Uinta Mountains owe their existence to this deep-seated structure. The Uintas are an inverted rift valley that formed along a much later reactivation of the structure. The rift, filled with lowdensity quartzite, was easy to push up during the Laramide.38 The Paleozoic was a time of relative quiescence in the Red Desert, the Pacific a passive margin mantled by thin deposits of limestone, shale, and quartzite. The major mountain belts of this era were along the Atlantic and Gulf coasts, where continental collision with Europe, Africa, and South America was building an orogenic welt of Himalayan proportions. The Ancestral Rockies, foreland uplifts interior to the Gulf Coast Ouachita orogeny of Oklahoma and Texas, formed at this time, sweeping northwest across the Texas panhandle, New Mexico, and central Colorado and shedding a bit of red sediment northward into Wyoming. In the early Mesozoic, the Red Desert lay in the trade wind belt, where northeasterly winds and arid conditions persisted for long periods of time. Sand that had swept across the continent from the east during the Atlantic orogenies was reworked into dunes.39 As the Atlantic opened again and the east coast changed into the passive margin that it is today, subduction was initiated along the Pacific. This was just the beginning of a series of oro90

genic events that were to mold, twist, and contort the Pacific margin for the next 200 million years. For a long time though, all Wyoming ever knew of it was that sediment streamed eastward across her craton. The Mesozoic tectonic wave didn’t reach Wyoming until the late Cretaceous. By then continental drift had moved her north into the temperate belt, shifting the wind around to the southwest. Sediment continued to stream across Wyoming from the west, but now the prevailing westerlies brought volcanic ash as well.40 The Laramide and Lake Gosiute The first hint that there was a Laramide Revolution in Wyoming was that her craton started to subside.41 Its middle Mesozoic erstwhile erg—windswept sand dunes with little vegetation—probably at low elevation already, succumbed to a marine Cretaceous sea.42 Tides, sharks, and plesiosaurs swept and swam across Wyoming from the east while sediment streamed in from the rising mountains to the west. Deltas and beaches built, swamps expanding behind them. Dinosaurs. Coal. Fifteen million years into the Paleogene, Wyoming’s foldthrust belt continued to rise and advance to the east, but the hardened craton resisted.43 With thousands of kilometers of oceanic crust still to be consumed along the Pacific margin, something had to give. Tectonic forces shifted to discrete deep crustal fault systems, leading to the emergence of foreland uplifts well to the east. In the process, a Cretaceous ocean was traded for an Eocene inland sea, Lake Gosiute. Old, cold, and stable, cratons worldwide are no strangers to lakes and inland seas. Most of the world’s big lakes, modern and ancient, representing all major classes,44 seem to have an affinity for Archean crust. The glacial lakes that circle the Laurentian shield, the rift lakes of Africa, and the foreland landlocked seas of Asia, South America, and North America almost all lap onto Archean shores.45 Perhaps the biggest and longest-lived of them all lapped onto the Wyoming craton from the northwest in the

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 91

Mesoproterozoic.46 All of northwest Montana, northern Idaho, eastern Washington, southeastern British Columbia, and southwestern Alberta were inundated. The lake was so huge, in fact, that its western shoreline got rifted away in the Neoproterozoic. It now lies on another continent, lapping onto the Siberian craton.47 As Wyoming’s foreland uplifts continued to grow and evolve, so changed Lake Gosiute, shrinking and swelling in response to the subtle interplay of climate change,48 tectonics,49 and evolving sediment supply.50 It swelled and shrank three times, the first two while retaining a connection to the Atlantic,51 a fishy Lagerstätten in the making.52 During its third phase it filled and spilled again, but this time to the south.53 Green River Formation lake beds were deposited continuously in the center of the basin; thus, nearly the entire thickness of White Mountain overlooking Rock Springs consists of lacustrine strata— cyclic oil shale laminite, evaporite, and ledge-forming stromatolite.54 From the center of Lake Gosiute’s depocenter toward its margins, the Green River Formation splits into three members: the Luman, Tipton, and Laney, each shooting a tongue of faded gray and cream shale slope and limestone ledge into the brightly colored, easily eroded, variegated alluvial sandy mudstone of the enveloping Wasatch Formation. The Wasatch is the badland maker of the Red Desert, sort of a Chinle of the north choked with rounded pebbles, petrified wood, and vertebrate fossils.55 The interfingering Wasatch badland and Green River ledge landforms dictate the shape of the desert’s landscape. The hogback rims circling the margins of the synclinal Green River and Washakie basins, and the stairstep “Escalante” rims of the Steamboat–Honeycomb Buttes region all owe their existence to Lake Gosiute’s complex Eocene history. “Minerals containing the sulfate radical are extremely rare,” writes H. Bradley in the abstract for his authoritative paper on the Green River Formation.56 Bradley’s understated pun, intentional or not, is appropriate. Practically everything about the Green River Formation is either radical, extreme, or rare. In between its second and third expansions, Lake Gosiute got radical. Changing

from spill and fill to sit and stew, it shrunk and stayed small for a prolonged period of time, resulting in a hypersaline unit known as the Wilkins Peak Member, whose evaporites are more than a little odd. Extreme is more like it.57 The barren, sterile granite and quartzite cores of the two main mountain ranges that were furnishing most of the water to the basin produced world-class deposits of soda ash, trona, and shortite.58 The kelp burners could stop their dirty business.59 The last filling of Lake Gosiute was also accompanied by a major influx of volcaniclastic sediment derived from the Absaroka volcanic field to the north.60 When Lake Gosiute filled for the final time it was this sediment that killed it.61 The blue-green conglomeratic deposits, punctuated by swatches of vermilion sandstone and mudstone with sparse, thin bands of chalk-white volcanic ash, crop out mostly toward the middle of the Green River and Washakie basins, where they form breached dip-slope rims that face inward.62 The Blue, Adobe Town, Skull, and Prehistoric rims are of this type. The Neogene In the Red Desert, the Neogene is more hiatus than record, and the formations that record it are complex, sparsely preserved, and poorly understood. In the Oligocene, the last epoch before the Neogene period, a pulse of post-Laramide uplift migrated across Wyoming. This event, referred to as the Wyomide by Jim Steidtmann,63 probably represents what happened to Wyoming in response to the subduction of the East Pacific rise. Sedimentary rocks of this age are preserved in widely separated belts flanking the ranges that were uplifted, the Wind River–Granite mountains to the north, the Uintas to the south, and the Sierra Madre–Rawlins Uplift to the east. For the northern and southern ranges, uplift was recorded by fans of boulder conglomerate that built out and around both ranges,64 but in the Miocene, the ranges collapsed, tilting back gently along the same faults that had accommodated the uplift. The resulting grabens and half grabens filled

Geology of the Red Desert

91

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 92

gradually with fine-grained alluvial deposits: sandstone, siltstone, and mudstone.65 The low-gradient, slow-moving, low-volume streams were succeeded by playa lake beds interleaved with eolian sandstone.66 A Switch of Oceanic Allegiance Currently, two of the three Red Desert basins—the Green River and the Washakie — contribute their waters to the Green. The third, the Great Divide basin, contributes only to itself.67 Three seas vie for tribute from the Red Desert. By chance, the smallest, the Sea of Cortez, flies its banner there now. But its hold is tenuous. Subject to the whim of evolution,68 climate change, and plate tectonics, emissaries from the North Pacific and the Gulf of Mexico are eager to claim her water, both remembering that for long intervals of the Cenozoic they held sway. The fish fossils of the Green River Formation and the current fish fauna of the Green River are drastically different. During the Paleogene, Lake Gosiute was populated with fish from the Atlantic,69 but the current fish fauna of the Green River are Pacific in origin. The Green River has been integrated with the Colorado since at least the middle Pleistocene, but because most of the Neogene is so poorly preserved in Wyoming, the timing of the switch of oceanic allegiance is not well understood. When Lake Gosiute filled with sediment at the end of the Eocene it appears to have lost its connection with waters of the Atlantic, a disconnect that was temporary according to some 70 but may have persisted until today. An important debate is brewing, and the implications are not trivial. Here the story of the lower Colorado and its greatest tributary, the Green, must be considered together. The lower Colorado’s Neogene history is very complete, the depth of its time mirrored by the depth of its canyon. The upper Colorado and the lower Colorado have been integrated since the Grand Canyon was cut during the Miocene-Pliocene transition between 5.5 and 5.2 million years ago.71 These conclusions are nearly universally accepted, a rare thing in geology. A de92

bate continues, though. How did the river do its work? Headward erosion driven by recent uplift of the Colorado Plateau, say some.72 Spillway cut as Lake Bidahochi poured through a gap in the Kaibab Uplift, say others.73 Although the course of the ancestral Green River in Wyoming during most of the Neogene is unclear, it is known that during the early Pleistocene the Green River, at that time flowing across the eastern Red Desert, was captured.74 The Canyon of Lodore and Split Rock Gorge in the eastern Uintas, along with the Flaming Gorge, are all remarkably young, having been cut by the Green River in the last 620,000 years.75 Before that time the Green River was depositing gravels at Creston Junction near the eastern edge of the Red Desert. That this happened sometime since 620 ka is known because the gravel overlies a distinctive volcanic ash from Yellowstone whose age is precisely known.76 Furthermore, the gravel contains pebbles that could have come only from the west.77 The question is, to what place from there? The prevailing view is championed by Hansen, who concludes that it must have been to the Gulf of Mexico via the Platte and, as evidence, plots a series of river gravel terrace deposits that occur, conveniently, along a nearly straight line between Creston Junction and the town of Green River, a distance of approximately 150 kilometers. What Hansen fails to report, however, is that there are many other gravel deposits in the eastern Red Desert.78 Hansen’s solution creates more problems than it solves. Three largely unsubstantiated conclusions regarding other aspects of the area’s natural history must be accepted in order to satisfy the conditions of a Pleistocene Green-Platte connection. First, since the gravels occur in terraces that gain elevation to the east, Hansen is forced to conclude that a differential uplift of nearly 300 meters has occurred very recently in the eastern Red Desert. There is no independent evidence for this.79 Second, a Green-Platte connection in the Pleistocene should have resulted in a mingling of its fishes. There is, in fact, very little connection, a problem that Hansen sidesteps by arguing that capture must have occurred during an intense glaciation or dur-

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 93

ing some other equally brutal climatic event that wiped out all of the Platte’s headwater fishes. Great Basin ichthyologist Gerald Smith doesn’t buy this fish story. He states that Pleistocene glaciations elsewhere in the Rocky Mountains are not known to have resulted in these kinds of whole-scale exterminations, and in fact, he continues, the cooler, wetter conditions would have been favorable for fish survival.80 The third problem is that just as the Green has virtually no connection with the Platte, it has a very strong connection with freshwater fishes of the North Pacific. Hansen concludes that the shared lineages are the result of recent, post-glacial, headwater captures between the Green, Snake, and Bear rivers.81 The generic evidence, however, strongly indicates that North Pacific interlopers have been in the Green River system for millions of years.82 Based on the fish evidence, it is very unlikely that either the upper Green or the upper Colorado was connected with the Platte since the Miocene.83 The Green-Snake connection, however, is very strong,84 the fish evidence indicating that the Green River basin contributed its waters to the North Pacific during long periods of the Neogene via a connection through the ancient Snake-Sacramento river system.85 Based on scientific evidence, a cryptic river remarkably similar geographically to the mythical Buenaventura might have existed after all. Humans just missed it by 5 to 10 million years. An Alternative Exit Other river gravel deposits besides the one at Creston Junction have been identified in the Great Divide basin.86 They form terraces capping high points, rims, and buttes and can be grouped into discrete sets, based on similarity in elevation, that step down to the west toward the middle of the basin. The terraces’ step-down-to-the-west geometry suggests that they were deposited by a south-flowing river. Also, very importantly, they are all high enough to be upstream deposits of a river that could have conceivably exited the Great Divide basin through

its local low point at Coal Gulch (2,050 meters), about 50 kilometers to the southeast. Furthermore, the drainage immediately to the south of this point —Muddy Creek, a tiny tributary of the Yampa via the Little Snake—is underfit, meaning that its valley is much bigger than its stream.87 The underfit valley of Muddy Creek includes a suspicious wind gap near Dad, about 35 kilometers southeast of the divide, called Blue Gap (2,010 meters), and just to the south a gravel terrace contains limestone cobbles that could not have come from the east (Sierra Madre) or the south. A Pleistocene upper Green River exiting the Great Divide basin to the south and debouching into the upper Yampa, and ultimately into the lower Green, is a better fit for the available geologic and biologic evidence. If the Green captured herself,88 then there would be no reason to expect a Green-Platte connection. There would also be no need to invoke rapid recent uplift in an area that bears no evidence of such events or to tie together drainage basins whose fish fauna argue strongly against any connection. So what was the Green River up to prior to the Pleistocene if it wasn’t feeding some Atlantic drainage? From the eastern edge of the Great Divide basin, the only conceivable route to the Pacific would have been to the south, and if it went that way, it’s hard to imagine a scenario in which it would not have integrated with the upper Colorado River of western Colorado, a stream that was, in the Miocene, already associated with a well-integrated canyon superposed across three major ranges.89 Considering that spillway cutting, caused by overflow of Lake Bidahochi, is the most feasible explanation for the formation of the Grand Canyon begs the question.90 Where did all the water come from? This question directs attention to one of the objections that has been raised to the fill-and-spill hypothesis for the formation of the Grand Canyon. The objection is that Lake Bidahochi, based on the sedimentology of its lake beds, seems never to have had a major inlet,91 and in this way it appears to have been an upland lake much like Lake Gosiute during its early stages. There is, however, a good explanation for the seemingly improb-

Geology of the Red Desert

93

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 94

able combination of lots of water, enough to fill the lake to beyond capacity, and lack of sediment. Suppose Lake Bidahochi filled so rapidly that it spilled before it had a chance to silt up? To satisfy this supposition, a sudden addition of water from upstream seems necessary: simply put, an upstream capture. The ancestral Green of Wyoming, with its good lower Colorado fish connection, seems then to be the only reasonable place to look for the additional water. If it was captured, then the question has to be asked, from where? The Snake is the most likely candidate. As it turns out, the current drainage divide between the Snake and the Green, in the northwestern corner of the Green River basin, is a headwater region that was undergoing three major tectonic adjustments at the end of the Miocene. First, the Gros Ventre Range was being thrust to the southwest.92 Second, there was renewed thrusting on the Wind River thrust.93 Third, the Yellowstone hot-spot bulge was in east-central Idaho at that time in an ideal position for it to interfere with the upper Snake drainage basin.94 These events might have caused the greater Green River basin, previously paying tribute to the upper Snake via a connection through the Hoback River, or possibly the Bear River,95 to be captured by the upper Colorado. Capture is envisioned to have involved a southeast-directed drainage that paralleled the thrust front of the Wind River Range.96 Miocene-Pliocene tectonically driven capture of the northern Green River basin by the upper Colorado would have set off a chain reaction, sending a pulse of hydrologic energy down the new river system, in turn causing Lake Bidahochi to rapidly fill, spill, and cut the Grand Canyon. The Leucite Hills The Leucite Hills lavas are weirdos, a fact clearly understood and marveled at by early igneous petrologists.97 The hills are a series of Pliocene-Pleistocene volcanic buttes and flows clustered along the continental divide on the northern edge of the Rock Springs Uplift. The bizarre, mantle-derived, phlogopite-phenocryst phono94

lite lavas are the world’s most potassic igneous rock.98 The lavas were erupted at roughly the same time as the massive eruptions that formed the calderas of Yellowstone, Long Valley, and Valles— by volumetric comparison, a mere drop in the bucket, but in terms of geologic significance a dipperful at least. Most of the lavas occur to the south of the gap at Steamboat Mountain, and in most cases each flow is attached to a vent.99 Two of the four exceptions, North and South Table mountains, both of which are second only to Steamboat Mountain in elevation, are conspicuous. They, along with Steamboat Mountain, form the gap, a 400meter-deep gorge on the continental divide (2,228 meters), currently a wind gap through which spills eolian sand from the Killpecker dune field. The possibility that it may have been cut by a major river, possibly the ancestral Green, is intriguing. Lava flows, because they can be precisely dated, and because they tend to flow into low areas near local base level, provide an excellent record of landscape evolution. Recent high-precision 40 Ar/39Ar analysis provides dates of all principal exposures of the Leucite Hills lavas and makes important conclusions regarding the origin of tectonic plateaus and the composition of the Red Desert’s nether mantle regions.100 The oldest lavas — at South Table Mountain and Pilot Butte—are between 2.5 and 3.0 million years old, respectively. The highest lava, at Steamboat Mountain, erupted 1.7 million years ago. The lowest, at Black Rock, is almost exactly 1 million years younger.101 Two significantly different incision rates for the area were calculated.102 The area to the south of Steamboat Gap and Black Rock Flat indicated modest, desertlike erosion rates of barely more than a centimeter every hundred years, but to the north the incision rate is nearly double. This was recognized as a problem but was dismissed by stating that the two principal flows that were used to determine the rate in the northern area, Steamboat Mountain and Black Rock, must be separated by a fault. However, a quick glance at any of the excellent geologic maps of the area shows that there are no faults.103 The abrupt northward increase in erosion rate at Steamboat Gap might be better explained by rapid erosion due to a major

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 95

throughgoing river. The cone at Black Rock might have been emplaced in the valley of such a river, and at 2,133 meters the rock is high enough for it to have been in, or near, the upstream course of a river that deposited almost all of the gravel terraces of the eastern Great Divide basin. Sand Dunes Active dunes in the Red Desert occur in two main areas; the Killpecker field in the west and the Ferris field in the east. Killpecker dune sand contains, on average, 9.5 percent heavy minerals,104 a very high proportion, especially in comparison with the Ferris field, where sand contains less than half of a percent nonopaque heavy minerals.105 The only source in the area that can deliver such immature sand is the glacial outwash fan of the combined Big and Little Sandy rivers, which conveniently lie just upwind of the dune field. This unconsolidated sediment was originally interpreted to be the source of the Killpecker Dunes,106 an interpretation later dismissed based primarily on the heavy minerals.107 Killpecker and Ferris dune sand is immature,108 and it was ultimately derived from erosion of the Wind River Mountains. In this regard, the speculative suggestion that the Killpecker dune field and the string of dunes that trail from it across the Red Desert might represent the ghost of a major river is conceivable. The notion that a gravel lag might develop in a desert region in response to fluvial erosion and preferential removal of a large volume of sand-size and finer sedimentary particles (thereby leaving behind a gravel lag) can also be applied to the notion of a sand lag left behind by another important desert process—wind. Troy Péwé reveals some important facts about desert dust.109 Five hundred million tons of dust per annum are eroded from the world’s deserts,110 and as much as half of the total sediment delivered from the continents to the oceans each year is probably achieved by the wind.111 The monstrous haboobs (dust storms) of the world’s deserts therefore have an enormous impact on global geology,112 especially since deposits of this dust, in the form of

water-saturated, deep-sea pelagic sediment, have been providing the world’s subduction zones and magmatic arcs with its chief supply of fluxing agent for billions of years. The mysterious sand sea of the Red Desert might in fact represent the wind-deflated sandy alluvial remains of the ancestral Green and Sweetwater rivers, still after half a million years moving slowly downwind and escaping slowly through the wind gap between Junk Hill and Bear Mountain at the east end of the Ferris Mountains.113 The clay and dust have long since blown away, most deposited on the bottom of the Atlantic Ocean. It’ll take millions of years for the sand to get there. But the desert has time. Summation The Paleogene Green to the Neogene Green was a switch of oceanic allegiances. Big Lake Gosiute filled up and poured into the Atlantic because the western cordillera was just too massive and impenetrable. It was like an Andes, so big that a river as bad-ass as the Amazon couldn’t even wind through, and flowed east because it had to. Things changed. The tectonic wave that had rocked the Pacific margin so hard in the middle Mesozoic was finally being felt in the interior. Weird foreland uplifts popping up inboard complicated the issue, making it hard for rivers to stay their course. But they did, for a while anyway. Then it came. Not the uplifts, not the hard-margined faults that broke her and forced her rivers to flow into new, uncharted channels. It was the bulge, the deep-seated inflation of the continent that finally did it. The head of Midgard’s serpent subducted, shoving first the Colorado Plateau, and then its northern cousin the Red Desert, up, riding high, biding time, letting their rise be felt by the slow tide and trade of tribute, river to river, basin to basin, sea to sea.

Geology of the Red Desert

95

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 96

Abbreviations aapg ajs gca gsa gsw ofr sepm sp usgs uw wga wgs

96

American Association of Petroleum Geologists American Journal of Science Grand Canyon Association Geological Society of America Geological Survey of Wyoming Open-File Report Society of Economic Paleontologists and Mineralogists Special Paper United States Geological Survey University of Wyoming Wyoming Geological Association Wyoming State Geological Survey

Notes 1. Love (1961) formally defines the basins. 2. Jason Lillegraven (1993) writes, “In reference to Paleogene time, the geologic record of Wyoming is better understood than any other comparably sized tract of Earth’s dry land.” 3. Jack Repcheck’s biography (2003) of the father of geology, James Hutton, starts with this from Stephen Jay Gould: “He burst the boundaries of time, thereby establishing geology’s most distinctive and transforming contribution to human thought—Deep Time.” 4. Samuel H. Knight (1955). 5. Darwin (1859). 6. Knight (1955). 7. WWII-driven technology. 8. Add 8,000 kilometers more for its only arm, which projects north from the serpent’s belly in the Indian Ocean. The arm forms the Afar Triangle and the great East Africa Rift before it gets swallowed up by the Tethyan orogen. Oceanographer Joellen Russell uses the Midgardian serpent analogy for oceanic currents. 9. A 10-centimeter-wide sliver of crust, 10 kilometers thick and 50,000 kilometers long, is 50 cubic kilometers. By comparison, the concrete in Hoover Dam, enough to pour out a four-lane highway stretching coast to coast, would have to be multiplied 275 times to fill a single cubic kilometer. Fifty of them are the equivalent of nearly 14,000 Hoover Dams. 10. The age of the world’s midocean ridge is at least 2 Ga. Slivers of oceanic crust, also known as ophiolites, born of it have been thrust onto the continents in ancient mountain belts. Its age could therefore be even older, possibly in excess of 4 Ga. 11. The plate tectonic process in which one plate is overridden by another. 12. Up until this time, the shrinking earth theory offered a reasonable explanation for why the world had so many convergent mountain belts. The wrinkles on the skin of a shrunken apple provide the analogy (Hineline, 2005). Since the structure and composition of the earth’s nether mantle region was completely unknown, the notion that it might be in a dynamic state of contraction was seriously considered (after all, lots of things were known to contract when they cooled, from loaves of bread to ingots of iron). Since the crust was already cool and rigid, like the skin of an apple, it had to deform by crumpling instead. See also Kious and Tilling (1996). 13. The process of mountain building, or orogenesis (from the Greek Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 97

oros, meaning “mountain”), results in orogenic belts, or, simply, an orogeny. 14. With Mount Everest at 9 kilometers high, and an active belt 2,500 kilometers wide, including the Tibetan Plateau and Himalaya. 15. The length is 42,000 kilometers if you picture the earth as a baseball, cut stitches, and exclude a weird Sandwichian Islands twitch in the South Atlantic. 16. Continents don’t like to lose. As irrefutable evidence of this, the world’s youngest oceanic crust, located in the northwest Pacific, is only 140 Ma, whereas the cratonic cores of each continent are at least 2,700 Ma, nearly ten times and then two times again as old. 17. Multiple folding events in the internides, the metamorphosed core of fold belts, result in strongly foliated micaceous rocks, slate and schist, whose internal structure resembles that of the multiply folded layers of a samurai’s sword. 18. When Salt and Wyoming Range thrusts were last active, approximately 50 Ma, the Pacific was several hundred kilometers closer. Up to 100 percent extension has occurred across the Great Basin since the middle Cenozoic, approximately 30 Ma. 19. Erslev (1986) argues convincingly that these have to feed into a midcrustal decollement, a zone of ungluing, along which thrusting, driven by Pacific convergence, is transmitted deep into the continent. 20. The Red Desert is renowned for its sodic evaporites (Bradley and Eugster 1969). South Pass’s Archean iron formation, or taconite, recalls a time when earth’s atmosphere was void of free oxygen. Wyomingite, the main type of lava found in the Leucite Hills, was named by Cross (1897). 21. Foreland uplifts are conspicuously composed of large monotonous bodies of low-density rock, granite and quartzite especially. Such are the Red Desert’s two biggest ranges, the Winds and the Uintas, respectively. 22. The sight of these impenetrable mountains gravely disappointed the young Alexander Mackenzie in 1789 as he descended the great river that was eventually named for him. In fact he had driven one of the final nails into the coffin of the great search for the Northwest Passage. Four years later, he crossed them and reached the Pacific (DeVoto 1952). 23. The fantastical inland western sea was supposed to feed a westflowing river, the Buenaventura, that found the Pacific near Sacramento (Wheat 1958). 24. Bolton (1951); DeVoto (1952). 25. Goetzmann (1993). 26. D. L. Morgan (1953); D. L. Morgan and Wheat (1954). 27. Gilbert (1983). 28. Fremont (1973).

29. The Archean, first eon of earth history, ended 2.5 Ga. The geologic term “sial” is a periodic-table acronym that refers to the stuff of continents, rocks enriched in silica (SiO2) and alumina (Al2O3). 30. Condie (1981). 31. The thermochronologic work of Danny MacPhee et al. (2005) in the Rwenzori Mountains, Congo, East Africa Rift, shows that granites higher than 5 kilometers on Mount Margherita are Archean. Uplift has been so rapid, however, that it is possible the Red Desert may have held the title of world’s highest Archean crust sometime during humankind’s not too distant past. 32. Condie (1981); Martin (1994). Antarctica may be an exception here, but when the ice melts, it is likely that a mate for the Napier Complex will be revealed. 33. Hoffman (1988). 34. Chamberlain et al. (2003). 35. Most of the large blocks of jade that used to litter Crooks Mountain have been collected (Love 1960). Jadeite is one of a group of minerals that attend high-pressure, low-temperature metamorphism, a hallmark of subduction zones (Miyashiro 1973). 36. The Cheyenne mobile belt is approximately 1,900 Ma (Houston et al. 1989). The strike is a line produced by the intersection of any dipping planar surface with the horizontal. The trend is a purely linear feature. A linear valley, for example, has a trend. See DeVoto (1952, 405) for an example of the proper use of these terms by a nongeologist, in his case in reference to, of all places, the northern Appalachians. 37. The Independence Mountain thrust, an important young structure of this orientation, but curiously opposite dip, parallels the state line where the North Platte enters Wyoming (Blackstone 1983). 38. Long thought to be Mesoproterozoic, the Uinta quartzites are Neoproterozoic (Dehler et al. 2005). 39. Dickinson and Gehrels (2003). 40. The Red Desert’s uranium resources may owe their existence to this. The source of the uranium is debated (the Granite Mountains are enriched in this element to begin with), but one of the most likely explanations is that it blew there. Tell that to a Jeffrey City miner after a day of fighting midwinter Crooks Gap katabatic winds, and he just might believe you. Uranium, despite its high density, is also a large-ion lithophile element, meaning that when the crust melts, it is one of the first elements to enter the magma, and also one of the last to crystallize. For this reason, it concentrates in silicic, gas-rich magmas. Since these rocks tend to erupt catastrophically, fine-grained, uranium-enriched ash was spread liberally and repeatedly across the Red Desert for approxi-

Geology of the Red Desert

97

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 98

mately 150 million years. The uranium, because it is easily soluble in oxidized water, was then washed out of the volcanic ash and transported great distances in water-saturated sediments and sedimentary rocks. At least four kinds of uranium deposits have been documented in the Red Desert (Love 1964; Pipiringos 1961; Masursky 1962). Three of them— the lake-bed phosphates of the Wilkins Peak Member, the uraniferous coal seams of the Wasatch Formation, and the world-renowned desert evaporite schroeckingerite deposits of the Lost Creek area—are of interest only to academics and collectors. The fourth type is economic: rollfront deposits of the northern Great Divide Basin. Oxidized formation waters working their way toward the centers of their basin abruptly precipitate their uranium whenever they encounter reducing conditions; coal and organic-rich sediment do the trick. The deposits resemble rolls and occur in elongate fronts. 41. Geologists used to use the term “revolution” in place of “orogeny.” 42. An erg is a sea of sand dunes, erstwhile, because the long-lived Permian-Jurassic erg’s northern boundary fluctuated across the Red Desert region. 43. The Wyoming salient, the frontal arc of the fold-thrust belt in Wyoming, is constrained by the rigid western extremities of the Wind Rivers and Uintas. For timing of orogeny, see the review by Royse (1993). 44. Cohen (2003, Table 2.1). 45. The Caspian Sea is the only major exception, but it deserves mention since its volume exceeds that of all the other major lakes of the world combined (ibid., Table 2.2). 46. The Belt-Purcell Supergroup. See Walcott (1914) and Winston (1986). 47. Sears and Price (1978, 2003). 48. Bradley and Eugster (1969); Roehler (1993); Oviatt (1997). 49. Surdam and Stanley (1980). 50. Pietras et al. (2003); Carroll et al. (2006). 51. Bradley (1964). 52. “Lagerstätten” freely translated by Gould (1989) is “mother lode.” 53. Bradley (1929). 54. Paper-thin laminae of oil shale, evidence of annual algal blooms, were deposited during high-stands of the lake. The stromatolites, limestone banks deposited in the form of hardened algal mats, covered wide swaths of shallow lake bed inward of the strand line, which would migrate back and forth great distances across the flats as the lake swelled and contracted during each phase. See Eugster and Surdam (1973); Eugster and Hardie (1975); Surdam and Wolfbauer (1975). 55. The Chinle Formation is perhaps the Rocky Mountains’ most dis98

tinctive badland maker. The petrified forest in Arizona and the Abiquiu– Ghost Ranch area of New Mexico are prime examples. 56. Bradley (1964). 57. Bradley and Eugster (1969) report that the conditions of evaporite sedimentation were so extreme that a slew of authigenic oddballs formed. Until described in the Green River Formation, they were thought to crystallize only at what the authors describe as “rather high temperatures.” 58. Ibid. Shortite, a sodium carbonate, is found only in the Green River Formation. 59. The production of glass, an essential item for the industrial age, requires soda ash. The British Isles’ lack of this material forced the industry to burn kelp gathered by beachcombers (Kostick 1998). 60. Surdam and Stanley (1979, 1980). 61. That the Absaroka was, and still is, a great supplier of sediment is evident at the mouth of the Mississippi; a lion’s share of the sand currently delivered to the delta probably comes from this source (Reiners et al. 2005). 62. Conglomerate is a sedimentary rock composed predominantly of particles coarser than granule sand (bigger than 4 millimeters). Wentworth (1922) defines these particles as pebbles (4– 64 millimeters), cobbles (64 to 256 millimeters), and boulders (greater than 256 millimeters). 63. These uplifts have exerted a strong influence on Wyoming’s geomorphology (Mears 1993). 64. Wagon Bed, Ice Point, and Crooks Gap conglomerates in the north (Love 1970); the Bishop Conglomerate in the Uintas (Hansen 1984, 1986). 65. Grabens and half grabens are valleys bounded by normal faults. To the north, in the Granite Mountains the White River, Split Rock, South Pass, Circle Bar, and Moonstone formations were deposited (Love 1970; Steidtmann and Middleton 1986; Steidtmann et al. 1986, Flanagan and Montagne 1993; Sutherland and Hausel 2005). To the south, in the Uintas, an equally complex sequence is referred to by the name of its most prominent half-graben valley, Browns Park (Hansen 1984, 1986). Farther east, the Brown’s Park consists of very thick eolian sandstone deposits (Buffler 2003). 66. Eolian, from the Greek Eolus, god of the winds, meaning windblown sediment. 67. The Great Divide Basin is neither the highest nor the largest of its kind in North America, much less along the cordillera of the Americas. The Plains of San Agustín in central New Mexico are slightly smaller but also slightly higher. The Mimbres Basin and its relatives in southern New Mexico–Chihuahua, stretching from just west of Las Cruces to the Ari-

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 99

zona border, is much larger. 68. Castor canadensis (beaver) isn’t the only species capable of redirecting rivers. Two of the Red Desert’s drainage divides have been lowered significantly by human activity. The U.S. Highway 287 roadcut south of Muddy Gap suddenly made it the third-lowest point on the continental divide (2,055 meters). A railroad cut just west of Kemmerer has lowered the Bear River–Green River divide by 40 meters. 69. Grande (1984). 70. Hansen (1985). 71. Lucchitta and Jeanne (2001); Spencer et al. (2001). 72. Lucchitta (1979); Lucchitta and Jeanne (2001). 73. Lake Bidahochi, whose fish fauna is linked to the Colorado-SnakeSacramento systems (Spencer, Smith, and Dowling, in review), inundated much of northeast Arizona in the late Miocene, just before the Grand Canyon formed. That it cut the canyon by spilling over its dam is an interpretation supported by a number of authors: Blackwelder (1934), Meek and Douglas (2001), Scarborough (2001), Spencer and Pearthree (2001). 74. Hansen (1985). Regarding the use of “captured” versus “pirated,” which has come to identify partisans of two camps arguing a scientific point, see discussion by Pederson (2001). Lane (1899) championed groundwater sapping as the leading cause of stream capture. Crosby (1937) argued that surface flow was the principal agent of stream piracy. 75. Hansen (1985). 76. The ash called Lava Creek B was correlated based on mapping by Sanders (1975), sampling by Dave Love, and analysis by Izett and Wilcox (1982). It is worthwhile noting that the ash’s age is based on correlation only—the variable iron content of the ash, to be exact. It has never been dated directly at this locality. Although Hansen (1985) states that the ash is overlain by the gravel, Izett and Wilcox (1982) state that the ash is “underlain by brown gravel.” 77. Distinctive pebbles in the deposit could have come only from the Wind River and Uinta mountains (Hansen 1985). 78. Ibid. 79. There should be structures, and the continental crust should be seismically active. Seismic activity here is about as quiet as it can be. No obvious structures are identified. 80. Gerald Smith in personal communication proudly recalls that his great-grandmother walked the Immigrant Trail over South Pass when she was seven. 81. Hansen (1985). 82. dna nucleotide sequences as a chronometer of branding points in

fish evolution (Spencer, Smith, and Dowling, in review; Estabrook et al., in review; Smith and Dowling, in review). 83. Of the thirty-nine fish species indigenous to the Platte and Green rivers, only one, the mountain sucker (Catostomus platyrhynchus), occurs naturally in both drainages (Spencer, Smith, and Dowling, in press). 84. The Green River shares seven of its thirteen species and seven of eight genera with the Snake River drainage (ibid). The Snake and Sacramento rivers, via a connection through the Klamath of northern California, have had strong generic connections since the late Miocene (Smith 1992). 85. G. R. Smith et al. (2000, 2002). 86. Wyant et al. (1956); Welder and McGreevy (1966); Pipiringos (1961); Masursky (1962); Love and Christiansen (1985); Grasso (1990); this report. 87. Other manifestly underfit valleys in the Red Desert, like the lower reaches of Bitter Creek and the entire course of Killpecker Creek, were also probably cut by much larger rivers before they too experienced capture. 88. In the immortal words of the Dude, “New shit has come to light, man. . . . She kidnapped herself” (Coen and Coen 1997). 89. Kirkham et al. (2001). 90. House et al. (2005). The filling of Lake Bidahochi isn’t the only problem that would be satisfied by the sudden introduction of large amounts of water to the river system. The Pliocene lakes of the lower Colorado are so large that, in order to fill them, flow of the Pliocene Colorado would have had to have been at least comparable to what it is today. Without contribution from the Green, it would have been virtually impossible to fill them (Spencer, Pearthree, and House, in review). 91. Dallegge et al. (2001). 92. Love (1956, 1960). 93. Naeser (1984) based on fission track thermochronology. 94. L. A. Morgan and McIntosh (2004, 2005). See also Armstrong et al. (1975); Pierce and Morgan (1992); R. B. Smith and Braile (1993); Beranek et al. (2006). 95. Supported by fish evidence (Hansen 1985; G. R. Smith et al. 2000, 2002). 96. Thrust loading would have depressed the foreland basin, causing a natural trough to develop in this area (e.g., Heller et al. 1988). Later, after thrusting stalled in the middle and late Pliocene, sediment would have filled the trough and forced drainage to flow away from the range, and in this way, the modern, south-flowing Green River drainage eventually developed.

Geology of the Red Desert

99

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 100

97. The Leucite Hills played an important role in the evolution of modern igneous petrology, championed by Whitman Cross and his colleagues, Joseph P. Iddings, Louis V. Pirsson, and Henry S. Washington as an example of why the classification of igneous rocks should be based on composition rather than texture. The normative calculation, a standard geochemical-petrologic technique for analyzing igneous rocks, is commonly referred to as the cipw norm, in honor of these four pioneering petrologists. 98. The top of the mantle, the next innermost layer of the earth, is in Wyoming about 60 kilometers deep. Phlogopite is a dark mica similar to but more potassic than biotite. Phenocrysts are minerals that crystallize in magma prior to eruption. Carmichael (1967) grants these rocks, renamed phonolite by modern convention, this distinction. The original name applied to most of the lava was “wyomingite” (Cross, 1897). 99. Ogden (1979). 100. Lange et al. (2000). 101. Based on its horizontal flow-foliation, Black Rock was interpreted as a flow by Lange et al. (2000), but Schultz and Cross (1912) and Ogden (1979) interpreted it as a plug, because the foliations at the base of the rock consistently dip inward. An alternative interpretation is that the rock is the remains of a cinder cone filled with lava. The dating of Black Rock is a problem. Lange et al. (2000) report two widely separated ages, a whole-rock date of approximately 0.8 Ma, and a phlogopite separate of approximately 1.33 Ma. They prefer the older age and use it to calculate their incision rates— even though the rock, despite being normal polarized, would have had to have cooled during the Matuyama Reversed Chron if the date they prefer is correct. This warrants a lengthy discussion in their paper. The younger date is preferred, mostly because it closely matches a date from an independent lab, an unpublished phlogopite separate date of 0.77 Ma ± 0.02 Ma graciously provided by McIntosh, Chesner, and Love. 102. Lange et al. (2000). 103. Shultz and Cross (1912), Smithson (1959), Johnston (1959), and Love and Christiansen (1985), just to name a few. 104. Ahlbrandt (1973, 1974) never discusses this fact. It was gleaned from the appendix of his dissertation (1973). 105. Gaylord (1982), also determined that the Ferris dune sand is arkosic, containing up to 34 percent feldspar. Ahlbrandt (1973, 1974) did not quantify the quartz-feldspar abundances of the Killpecker sand samples. He was so convinced that heavy mineral analysis alone would tell him what he needed to know that he did not bother to study the bulk of the sand in his samples (R. W. Jones, pers. comm.). Gaylord (1982) an100

alyzed one sample identified as Killpecker “tail” sand, which contained less than 2 percent nonopaque heavy minerals and about 10 percent feldspar. 106. Holmes and Moss (1955). 107. Based on heavy mineral assemblages, Ahlbrandt (1973, 1974) interpreted Green River Formation beach sandstone outcrops in the area to be the source. According to his own data tables (1973), however, this is an impossibility. Two analyses of the beach sandstone average 3– 4 percent heavy minerals (remember that his dune sand samples averaged 9.5 percent, nearly three times this amount). I know of no natural process that might concentrate the heavy minerals in such a way. Weathering of the sandstone could only have lowered the heavy mineral fraction, not increased it, and the process of eolian transport, likewise, could only have reduced it even more. Ahlbrandt’s single-minded approach reflects how popular science influences the course of science. Studying a sedimentary rock’s provenance by analyzing the nature of its heavy minerals was at the time a powerful and sexy new technique (see, for example, the fine papers by Denson and Pipiringos 1969 and Denson and Chisholm 1971). Techniques like this are wonderful, but when scientists focus too much on one aspect of petrology, they are apt to dismiss important observations that seem contrary to their way of seeing things. 108. Even though Ahlbrandt (1973) might disagree on whether it was first- or second-cycle sand, he would agree that the ultimate source was the Wind Rivers. Gaylord (1982) interpreted the source of the Ferris dune sand to be the Eocene Battle Spring Formation, an arkosic alluvial fan complex that was derived from erosion of the Granite Mountains, the eastern continuation of the Wind River Range. 109. Péwé (1981). 110. Peterson and Junge (1971). 111. Goudie (1978). 112. And on geopolitics as well (Bowden 2006). A haboob over Iranian airspace in April 1980 had an enormous impact on human history. Jimmy Carter’s presidency was swept away with it, and for good or ill, the world changed. 113. Gaylord (1982) calculates a much more rapid transport rate for the Ferris dune field, but his rate assumes current conditions. The sand may have been stored in vegetated dunes or within lacustrine basins for long periods of time during the Pleistocene.

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 101

References Cited Ahlbrandt, T. S. 1973. Sand dunes, geomorphology, and geology, Killpecker Creek area, northern Sweetwater County, Wyoming. Ph.D. diss., uw, Laramie. ———. 1974. The source of sand for the Killpecker sand dune field, southwestern Wyoming. Sedimentary Geology 11:39 –57. Armstrong, R. L., W. P. Leeman, and H. E. Malde. 1975. K-Ar dating, Quaternary and Neogene volcanic rocks of the Snake River Plain, Idaho. ajs 275:225–251. Beranek, L. P., P. K. Link, and C. M. Fanning. 2006. Miocene to Holocene landscape evolution of the western Snake River Plain region, Idaho: Using the shrimp detrital zircon provenance record to track eastward migration of the Yellowstone hotspot. gsa Bull. 118:1027–1050. Blackstone, D. L., Jr. 1983. Laramide compressional tectonics, southeastern Wyoming. Contributions to Geology, uw Department of Geology and Geophysics, 22:1–38. ———. 1991. Tectonic relationships of the southeastern Wind River Range, southwestern Sweetwater Uplift, and Rawlins Uplift, Wyoming. gsw, Report of Investigations 47. Blackwelder, E. 1934. Origin of the Colorado River. gsa Bull. 45:551–566. Bolton, H. 1951. Pageant in wilderness. Utah Historical Quarterly, Utah State Historical Society. Bowden, M. 2006. The Desert One debacle. Atlantic, vol. 297, 62 –77. Bowring, S. A., I. S. Williams, and W. Compston. 1989. 3.96 Ga gneisses from the Slave Province, NWT, Canada. Geology 17:971–975. Bradley, W. H. 1929. The varves and climate of the Green River epoch. usgs Professional Paper 158-E, 87–110. ———. 1964. Geology of the Green River Formation and associated Eocene rocks in southwestern Wyoming and adjacent parts of Colorado and Utah. Geological Survey Professional Paper 496-A. Bradley, W. H., and H. P. Eugster. 1969. Geochemistry and paleolimnology of the trona deposits and associated authigenic minerals of the Green River Formation of Wyoming. Geological Survey Professional Paper 496-B. Buffler, R. T. 2003. The Browns Park Formation in the Elkhead region, northwestern Colorado–south central Wyoming: Implications for late Cenozoic sedimentation. In R. G. Raynolds and R. M. Flores (eds.), Cenozoic systems of the Rocky Mountain region, 183–212. Rocky Mountain sepm.

Carroll, A. R., L. M. Chetel, and M. E. Smith. 2006. Feast to famine: Sediment supply control on Laramide basin fill. Geology 34:197– 200. Carmichael, I. S. E. 1967. The mineralogy and petrology of the volcanic rocks from the Leucite Hills, Wyoming. Contributions to Mineralogy and Petrology 15:24– 66. Chamberlain, K. R., C. D. Frost, and B. R. Frost. 2003. Early Archean to Mesoproterozoic evolution of the Wyoming Province: Archean origins to modern lithospheric architecture. Canadian Journal of Earth Sciences 40:1357–1374. Coen, E., and J. Coen. 1997. The Big Lebowski. Universal Pictures. Cohen, A. S. 2003. Paleolimnology: The history and evolution of lake systems. Oxford. Condie, K. C. 1981. Archean Greenstone belts. Elsevier, Amsterdam. Crosby, I. 1937. Methods of stream piracy. Journal of Geology 45:465– 486. Cross, C. W. 1897. Igneous rocks of the Leucite Hills and Pilot Butte, Wyoming. ajs 4:115–141. Dallegge, T. A., M. H. Ort, W. C. McIntosh, and M. E. Perkins. 2001. Age and depositional morphology of the Bidahochi Formation and implications for the ancestral Colorado River. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution. 47– 56. gca. Darwin, C. 1859. On the origin of species. John Murray, London. Dehler, C., P. K. Link, and C. M. Fanning. 2005. Mid-Proterozoic strata of northern Utah and southern Idaho: Dating and correlation of Uinta Mountain Group and Pocatello Formation (abstract). gsa Abstracts with Programs 37(7): 218. Denson, N. M., and W. A. Chisolm. 1971. Summary of mineralogic and lithologic characteristics of Tertiary sedimentary rocks in the Middle Rocky Mountains and the northern Great Plains. usgs Professional Paper 750-C, C117–C126. Denson, N. M., and G. N. Pipirongos. 1969. Stratigraphic implications of heavy-mineral studies of Paleocene and Eocene rocks of Wyoming. 21st annual field conference, wga, 9 –18. DeVoto, B. 1952. The course of empire. Riverside Press, Cambridge, MA. Dickinson, W. R., and G. E. Gehrels. 2003. U-Pb ages of detrital zircons from Permian and Jurassic eolian sandstones of the Colorado Plateau, USA: Paleogeographic implications. Sedimentary Geology 163:29 – 66. Erslev, E. A. 1986. Basement balancing of Rocky Mountain foreland uplifts. Geology 14:259 –262.

Geology of the Red Desert

101

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 102

Estabrook, G. F., G. R. Smith, and T. E. Dowling. In review. Body mass and temperature influence on evolutionary rate of mtDNA in North America Cyprinidae. Eugster, H. P., and R. C. Surdam. 1973. Depositional environment of the Green River Formation of Wyoming: A preliminary report. gsa Bull. 84:1115–1120. Eugster, H. P., and L. A. Hardie. 1975. Sedimentation in an ancient playalake complex: The Wilkins Peak Member of the Green River Formation of Wyoming. gsa Bull. 86:319 –334. Faure, G. 1977. Principles of isotope geology. John Wiley and Sons. Flanagan, K. M., and J. Montagne. 1993. Neogene stratigraphy and tectonics of Wyoming. In A. W. Snoke, J. R. Steidtmann, and S. M. Roberts (eds.), Geology of Wyoming (gsw Memoir 5), 572 – 607. Fremont, J. C. 1973. The expeditions of John Charles Fremont. Vol. 1. University of Illinois Press. Gaylord, D. R. 1982. Geologic history of the Ferris dune field, southcentral Wyoming. In R. W. Maars and K. E. Kolm (eds.), Interpretation of windflow characteristics from eolian landforms (gsa SP 192), 65– 82. Gilbert, B. 1983. Westering man. Atheneum, N.Y. Goetzmann, W. H. 1993. Exploration and empire. Texas State Historical Association, Austin. Goudie, A. S. 1978. Dust storms and their geomorphological implications. Journal of Arid Environments 1:291–311. Gould, S. J. 1989. Wonderful life. Norton. Grande, L. 1984. Paleontology of the Green River Formation, with a review of the fish fauna. gsw Bull. 63. Grasso, D. 1990. Recognition and paleogeography of Quaternary Lake Wamsutter (a proposed lake in Wyoming’s Great Divide basin) combining Landsat remote sensing and digital elevation modeling. Ph.D. diss., uw, Laramie. Hall, S. S. 2006. Marie Tharp, b. 1920: The contrary map maker. New York Times Magazine, December 31. Hansen, W. R. 1984. Post-Laramide tectonic history of the eastern Unita Mountains, Utah, Colorado, and Wyoming. Mountain Geologist 21:5–29. ———. 1985. Drainage development of the Green River basin in southeastern Wyoming and its bearing on fish biogeography, neotectonics, and paleoclimates. Mountain Geologist 22:192 –204. ———. 1986. Neogene tectonics and geomorphology of the eastern Unita Mountains in Utah, Colorado, and Wyoming. usgs Professional Paper 1356. 102

Heller, P. L., C. L. Angevine, and N S. Winslow. 1988. Two-phase stratigraphic model of foreland-basin sequences. Geology 16:501–504. Hineline, M. 2005. Getting to the core of the mountain theory. North County Times. Hintze, L. 1975. Geologic highway map of Utah. Brigham Young Geology Studies Special Publication 3. Hoffman, P. F. 1988. United Plates of America: The birth of a craton. Annual Review of Earth and Planetary Sciences 16:543– 603. Holmes, G. W., and J. H. Moss. 1955. Pleistocene geology of the southwestern Wind River Mountains. gsa Bull. 66:629 – 654. House, P. K., P. Pearthree, and M. Perkins. 2005. Tephrochronologic and stratigraphic constraints on the inception and early evolution of the Lower Colorado River support lacustrine overflow as the principal formative mechanism (abstract). gsa Abstracts with Programs 37(7): 110. Houston, R. S., E. M. Duebendorfer, K. E. Karlstrom, and W. R. Premo. 1989. A review of the geology and structure of the Cheyenne belt and Proterozoic rocks of southern Wyoming. In J A. Grambling, and B. J. Tewksbury (eds.), Proterozoic geology of the Southern Rocky Mountains (gsa SP 235), 1–12. Izett, G. A., and R. E. Wilcox. 1982. Map showing localities and inferred distributions of the Huckleberry Ridge, Mesa Falls, and Lava Creek ash beds (Pearlette family ash beds) of Pliocene and Pleistocene age in the western United States and southern Canada. usgs Miscellaneous Investigations Series Map I-1325. Johnston, R. H. 1959. Geology of the northern Leucite Hills, Sweetwater County, Wyoming. MS thesis, uw, Laramie. Kious, J., and R. J. Tilling. 1996. This dynamic earth: The story of plate tectonics. usgs. Kirkham, R. M., M. J. Kunk, B. Bryant, and R. K. Streufert. 2001. Constraints on timing and rates of late Cenozoic incision by the Colorado River in Glenwood Canyon, Colorado: A preliminary synopsis. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 113–116. gca. Knight, S. H. 1955. Review of the early geological explorations of the Green River basin area 1812 –1879. 10th annual field conference, wga, 10 –17. Kostick, D. S. 1998. The origin of the U.S. natural and synthetic soda ash industries. In J. R. Dyni and R. W. Jones (eds.), Proceedings of the First International Soda Ash Conference, wgs Public Information Circular 39, vol. 1, 11–33. Lane, A. 1899. A note on a method of stream capture. gsa Bull. 10:12 –14.

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 103

Lange, R. A., I. S. E. Carmichael, and C. M. Hall. 2000. 40Ar/39Ar chronology of the Leucite Hills, Wyoming: Eruption rates, erosion rates, and an evolving temperature structure of the underlying mantle. Earth and Planetary Science Letters 174:329 –340. Lillegraven, J. A. 1993. Correlation of Paleogene strata across Wyoming — a users’ guide. In A. W. Snoke, J. R. Steidtmann, and S. M. Roberts (eds.), Geology of Wyoming (gsw Memoir 5), 414– 477. Love, J. D. 1956. Summary of geologic history of Teton County, Wyoming, during late Cretaceous, Tertiary, and Quaternary times. In wga 11th Field Conference Guidebook, 140 –150. ———. 1960. Cenozoic sedimentation and crustal movement in Wyoming. ajs 258A:204–214. ———. 1961. Definition of Green River, Great Divide, and Washakie basins, southwestern Wyoming. aapg Bull. 45:1749 –1755. ———. 1964. Uraniferous phosphatic lake beds of Eocene age in intermontane basins of Wyoming and Utah. usgs Professional Paper 474E. ———. 1970. Cenozoic geology of the Granite Mountains area, central Wyoming. usgs Professional Paper 495-C. Love, J. D., and A. C. Christiansen. 1985. Geologic map of Wyoming. usgs. Lucchitta, I. 1979. Late Cenozoic uplift of the southwestern Colorado Plateau, and adjacent lower Colorado River region. Tectonophysics 61:63–95. Lucchitta, I., and R. A. Jeanne. 2001. Geomorphic features and processes of the Shivwits Plateau, Arizona, and the constraints on the age of the western Grand Canyon. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 65–70. gca. MacPhee, D., S. A. Bowring, and P. W. Reiners. 2005. Exhumation, riftflank uplift, and the thermal evolution of the Rwenzori Mountains determined by combined (U-Th)/He and U-Pb thermochronometry (abstract). gsa Abstracts with Programs 37(7): 449. Martin, H. 1994. The Archean grey gneiss and the genesis of continental crust. In K. C. Condie, ed., Archean crustal evolution, 205– 259. Elsevier, Amsterdam. Masursky, H. 1962. Uranium-bearing coal in the eastern part of the Red Desert area, Wyoming. usgs Bull. 1099B. Mears, B. M., Jr. 1993. Geomorphic history of Wyoming and high-level erosion surfaces. In A. W. Snoke, J. R. Steidtmann, and S. M. Roberts (eds.), Geology of Wyoming (gsw Memoir 5), 608 – 626.

Meek, N., and J. Douglass. 2001. Lake overflow: An alternative hypothesis for the Grand Canyon incision and development of the Colorado River. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 199 –206. gca. Miyashiro, A. 1973. Metamorphism and metamorphic belts. Allen and Unwin, London. Morgan, D. L. 1953. Jedediah Smith and the opening of the West. BobbsMerrill, Indianapolis. Morgan, D. L., and C. I. Wheat. 1954. Jedediah Smith and his maps of the American west. San Francisco Historical Society. Morgan, L. A., and W. C. McIntosh. 2004. 40Ar/39Ar ages of silicic volcanic rocks in the Heise volcanic field, eastern Snake River plain, Idaho. In W. W. Bonnichsen, C. M. White, and M. McCurry (eds.), Tectonic and magmatic evolution of the Snake River plain volcanic province (Idaho Geological Survey Bull. 30). ———. 2005. Timing and development of the Heise volcanic field, Snake River plain, Idaho, western USA. gsa Bull. 117:288 –306. Naeser, N. D. 1984. Fission-track ages from the Wagon Wheel no. 1, northern Green River basin, Wyoming, evidence for recent cooling. In B. E. Law, ed., Geological characteristics of low-permeability Upper Cretaceous and Lower Tertiary rocks in the Pinedale anticline area, Sublette County, Wyoming (usgs OFR 84-753), 66 –77. Ogden, P. R., Jr. 1979. The geology, major element geochemistry, and petrogenesis of the Leucite Hills volcanic rocks, Wyoming. Ph.D. diss., uw, Laramie. Oviatt, C. G. 1997. Lake Bonneville fluctuations and global climate change. Geology 25:155–158. Pederson, D. T. 2001. Stream piracy revisited: A groundwater sapping solution. gsa Today 11:4–10. Peterson, S. T., and C. E. Junge. 1971. Sources of particulate matter in the atmosphere. In W. H. Matthews, W. W. Kellop, and G. D. Robinson (eds.), Man’s inspection on the climate, 310 – 320. Cambridge, mit Press. Péwé, T. L. 1981. Desert dust: An overview. In T. L. Péwé (ed.,), Desert Dust: Origin, characteristics, and effect on man (gsa SP 186), 1–10. Pierce, K. L., and L. A. Morgan. 1992. The track of the Yellowstone hotspot: Volcanism, faulting, and uplift. In P. K. Link, M A. Kuntz, and L. B. Platt (eds.), Regional geology of eastern Idaho and western Wyoming (gsa Memoir 179), 1–53. Pietras, J. T., A. R. Carroll, and M. K. Rhodes. 2003. Lake basin response to tectonic drainage diversion: Eocene Green River Formation,

Geology of the Red Desert

103

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 104

Wyoming. Journal of Paleolimnology 30:115–125. Pipiringos, G. N. 1961. Uranium-bearing coal in the central part of the Great Divide basin. usgs Bull. 1099A. Reiners, P. W., I. H. Campbell, S. Nicolescu, C. M. Allen, J. K. Hourigan, J. I. Garver, J. M. Mattinson, and D. S. Cowan. 2005. (U-Th)/(HePb) double dating of detrital zircons. ajs 305:259 –311. Repcheck, J. 2003. The man who found time: James Hutton and the discovery of the earth’s antiquity. Perseus. Roberts, Sheila. 1989. Wyoming geomaps. wgs Educational Series 1. Roehler, H. W. 1993. Eocene climates, depositional environments, and geography, greater Green River basin, Wyoming, Utah, and Colorado. usgs Professional Paper 1506F. Royse, F., Jr. 1993. An overview of the geologic structure of the thrust belt in Wyoming, northern Utah, and eastern Idaho. In A. W. Snoke, J. R. Steidtmann, and S. M. Roberts, (eds.), Geology of Wyoming (gsw Memoir 5), 271–311. Sanders, R. B. 1975. Geologic map and coal resources of the Creston Junction Quadrangle, Carbon and Sweetwater Counties, Wyoming. usgs Coal Investigations Map C-73. Scarborough, R. 2001. Neogene development of the little Colorado River valley and eastern Grand Canyon: field evidence for an overtopping hypothesis. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 207–212. gca. Schultz, A. R., and W. Cross. 1912. Potash-bearing rocks of the Leucite Hills, Sweetwater County, Wyoming. usgs Bull. 512. Sears, J. W., and R. A. Price. 1978. The Siberian connection: A case for Precambrian separation of the North American and Siberian cratons. Geology 6:267–270. ———. 2003. Tightening the Siberian connection to western Laurentia. gsa Bull. 115:943–953. Smith, G. R. 1992. Phylogeny and biogeography of the Castomidae, freshwater fishes of North America and Asia. In R. L. Mayden, ed. Systematics, historical ecology, and North American freshwater fishes, 778 – 826. Stanford University Press, Stanford, Calif. Smith, G. R., and T. E. Dowling. In review. Paleohydrography and phylogeography of speckled dace, Rhinichthys osculus, in the Colorado River basin. gsa SP. Smith, G. R., N. Morgan, and E. Gustafson. 2000. Fishes of the MioPliocene Ringold Formation, Washington: Pliocene capture of the Snake River by the Columbia River. University of Michigan Papers on Paleontology 32. 104

Smith, G. R., T. E. Dowling, K. W. Gobalet, T. Lugaski, D. K. Shiozawa, and R. P. Evans. 2002. Biogeography and timing of evolutionary events among Great Basin fishes. In R. Hershler, D. B. Madsen, and D. R. Currey (eds.), Great Basin aquatic systems history (Smithsonian Contributions to Earth Sciences 33), 175–234. Smith, R. B., and L. W. Braile. 1993. Topographic signature, space-time evolution, and physical properties of the Yellowstone-Snake River Plain volcanic system: The Yellowstone hotspot. In A. W. Snoke, J. R. Steidtmann, and S. M. Roberts (eds.), Geology of Wyoming (gsw Memoir 5), 694–754. Smithson, S. B. 1959. Geology of the southern Leucite Hills, Sweetwater County, Wyoming. MS thesis, uw, Laramie. Spencer, J. E., L. Peters, W. C. McIntosh, and P. J. Patchett. 2001. 40Ar/39Ar geochronology of the Hualapai Limestone and Bouse Formation and implications for the age of the lower Colorado River. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 89 –92. gca. Spencer, J. E., and P. A. Pearthree. 2001. Headward erosion versus closedbasin spillover as alternative causes of Neogene capture of the ancestral Colorado River by the Gulf of California. In R. A. Young and E. E. Spamer (eds.), Colorado River: Origin and evolution, 215–222. gca. Spencer, J. E., P. A. Pearthree, and P. K. House. In review. Some constraints on the evolution of the latest Miocene to earliest Pliocene Bouse lake system and initiation of the lower Colorado River. gsa SP. Spencer, J. E., G. R. Smith, and T. E. Dowling. In review. Middle and late Cenozoic evolution of topography and late Cenozoic evolution of fish in the American southwest, and location of the continental divide. gsa SP. Steidtmann, J. R., and L. T. Middleton. 1986. Eocene-Pliocene stratigraphy along the southern margin of the Wind River Range, Wyoming: Revisions and implications from field and fissiontrack studies. Mountain Geologist 23:19 –25. Steidtmann, J. R., L. T. Middleton, R. J. Bottjer, K. E. Jackson, L. C. McGee, E. H. Southwell, and S. Lieblang. 1986. Geometry, distribution, and provenance of tectonic conglomerates along the southern margin of the Wind River Range, Wyoming. In J. A. Peterson, ed., Paleotectonics and sedimentation in the Rocky Mountain region, United States (aapg Memoir 41), 321–332. Surdam, R. C., and K O. Stanley. 1979. Lacustrine sedimentation during the culminating phase of Eocene Lake Gosiute, Wyoming

Charles Ferguson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 105

(Green River Formation). gsa Bull. 90:93–110. ———. 1980. Effects of changes in drainage-basin boundaries on sedimentation in Eocene Lakes Gosiute and Unita of Wyoming, Utah, and Colorado. Geology 8:135–139. Surdam, R. C., and C. A. Wolfbauer. 1975. Green River Formation, Wyoming: A playa-lake complex. gsa Bull. 86:335–345. Sutherland, W. M., and W. D. Hausel. 2005. Preliminary geologic map of the South Pass 1:1,000,000 quadrangle, Fremont and Sweetwater Counties, Wyoming. wsg ofr 2005-5. Tweto, O. 1979. Geologic map of Colorado. usgs. Walcott, C. D. 1914. Pre-Cambrian Algonkian algal flora. In Cambrian geology and paleontology III (Smithsonian Miscellaneous Collections, vol. 64, no. 2), 77–156. Welder, G. E., and L. J. McGreevy. 1966. Ground-water reconnaissance of the Great Divide and Washakie basins and some adjacent areas, southwestern Wyoming. usgs Hydrologic Investigations Atlas HA 219. Wentworth, C. K. 1992. A scale of grade and class terms for clastic sediments. Journal of Geology 30:377–392. Wheat, C. I. 1958. Mapping the Transmississippi West, vol. 2, From Lewis and Clark to Fremont. Institute of Historical Cartography, San Francisco. Winston, D. 1986. Sedimentology of the Ravalli Group, middle Belt carbonate and Missoula Group, Middle Proterozoic Belt Supergroup, Montana, Idaho and Washington. In S. M. Roberts (ed.), Belt Supergroup (Montana Bureau of Mines and Geology Special Publication 94), 85–124. Wyant, D. G., W. N. Sharp, and D. M. Sheridan. 1956. Reconnaissance study of uranium deposits in the Red Desert, Sweetwater County, Wyoming (usgs Bull. 1030-I), 237–308.

Geology of the Red Desert

105

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 106

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 107

2. Water in the Red Craig Thompson

ne is tempted to look at the harshness of the Red Desert and assume that catastrophic floods and cataclysmic winds shaped its rugged landscape. In fact, such rare meteorological events may shape a particular landform, but the real work of shaping an arid landscape is done incrementally and gradually. Seemingly endless tablelands, pancake-flat stacks of horizontally layered sediments, stretch from horizon to horizon. The sediments were first laid down in lakes or by rivers. Later, during more arid times, the giant lakes and rivers dried up. Smaller streams then sculpted the mesas and escarpments by channeling their moving water over the surface and incising deep valleys and steep canyons. In this desert, change must have occurred through infinitesimally small releases of energy over incomprehensibly large increments of time. The Green River is the western boundary of what we call the Red Desert, and it is the “Big Drain” of Red Desert hydrology. It drains an area five times larger than the closed Great Divide basin. The basin, ringed with red sediments from which the Red Desert takes its name, has no outlet, and water either evaporates or infiltrates into the groundwater. To understand water in the Red Desert is to understand great change. Imagine an ancient lake teeming with millions of fish whose transient chemical character still influences the water of

O

the Red Desert. Imagine a lush tropical forest with alligators and large turtles giving way to a dry desert now teeming with petrified wood and fossil evidence of ancient biodiversity. Lake Gosiute The Red Desert is a geographic basin of some 15,000 square miles filled with sediments from the Paleocene and later epochs. Between 10,000 and 20,000 feet of continuous layers of river and lake sediments accumulated in the basin.1 Gigantic meandering rivers that ran in the opposite direction of today’s Green River lofted clay, silt, and sand in their swift currents. The unceasing force of moving water rolled larger gravels, cobbles, and boulders along the streambed. These rivers served not only as slow-motion conveyor systems but also as elongated rock tumblers, hundreds of miles in length. First, the measured physics of weathering released sharp-edged angular rocks from their parent escarpments high in mountains that are today much closer to their progeny sediments. In time, the rocks were beveled, chamfered, rounded, and eventually polished on their leisurely journey to the basin. Along the way, chemical weathering helped to reduce their size, so that as they filled the Red Desert basin, the particles became the source material for sandstone, siltstone, and shale. Once in 107

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 108

the basin, the current slowed, and the sands, silts, and clays settled to the riverbed in that order. Building up higher and higher, the sediments caused the rivers to move sideways, made room for more sediments, and spread them throughout the basin. The mountain-building Laramide period gave definition to the Red Desert boundaries, and in the early Eocene, Lake Gosiute began to fill on top of the Paleocene sediments for the first time. Rimmed first by the Gros Ventre and the Wind River ranges to the north and the Wyoming Range to the west, and later by the Uinta Range to the south,2 the basin floor subsided in a gradual crustal warp that helped to provide depth and allowed successive layers to be deposited. These layers are called lacustrine sediments, meaning that they were deposited in inland lakes. Sediments that eroded from the peripheral mountains created wide alluvial plains and broad flat pediments. Lake Gosiute, at its maximum surface area of 15,000 square miles, flooded the Red Desert entirely and extended into Utah and Colorado.3 Lake Superior, containing 10 percent of the world’s freshwater, is the largest lake today—roughly two and a half times larger than Lake Gosiute. Lake Gosiute filled, spilled, rose, and fell, repeating this pattern through an astounding 4 million years. This Methuselah of lakes existed through geologic time, continental drift, and great climatic change. During the Eocene, the bays and shallow daughter lakes were alive with catfish, turtles, and crocodiles. Straightbilled flamingos plied the heavily forested and oft-flooded shorelines. While early researchers proposed that Lake Gosiute was a large, deep, and permanently stratified lake,4 the evidence is more consistent with a shallow “playa lake” model that expanded and contracted according to climatic fluctuations.5 During the wettest periods the lake reached its maximum extent, and during the driest periods probably evaporated completely.6 For a million years, Lake Gosiute was a freshwater lake with an outlet—presumably to the east through the Washakie basin.7 Then the climate changed and the lake shrank to two-thirds its former size. Re108

treating from its shoreline, Lake Gosiute became shallower and fell below the outlet. With no outlet and with constant evaporation during this arid period, the salinity increased. But the Lake Gosiute salinity had a different chemical character. Sometimes it was alkaline, a mixture of baking soda and sodium carbonate called trona; sometimes it was saline, like table salt. Constant evaporation caused the salinity to solidify in an efflorescent crust on the mudflat fringe of the lake—similar to the alkali flats of the Red Desert today. Like the Great Salt Lake and other closed-basin lakes, Lake Gosiute fluctuated in depth and size as it responded to changing climate and tectonic activity. For lakes located on flat playas, small changes in water-surface elevation result in large changes in lake surface area. For example, the Great Salt Lake tripled in surface area between 1963 to 1986, from 950 to 3,300 square miles, when the lake spilled over into the west desert. An elevation difference of less than nine feet caused the lake to increase by more than 44 percent in surface area. Lake Gosiute went through numerous regressive stages during which storm events and sheetwash transported the crusty alkaline salts into the lake’s center, making the remnant even more concentrated in salts. Before it became too saline, massive algae blooms gave the lake a biologically rich character. Photosynthetic algae fixed carbon from the water, and because the dissolved carbon dioxide was in equilibrium with the overlying atmosphere, the algae depleted the atmosphere of carbon dioxide (CO2). Like all plants, algae also respire CO2 and bloom and die. Settling to the bottom and decomposing in the presence of oxygen, the dead algae yielded CO2 back into the water. When the organic ooze combined with layers of fine sediments to seal oxygen out, the decomposition yielded methane. Plunge your hand into almost any lake-bottom sediments today and the smelly black mass that you retrieve is evidence of the same process. The smell is likely to be hydrogen sulfide — the anaerobic decomposition product of sulfur fixed by vegetation and then caught in the ooze. But algae and other aquatic plants require more than carbon

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 109

dioxide and water—they need inorganic nutrients like nitrate and phosphate. In Lake Gosiute the nitrate presumably came from the atmosphere through other nitrogen-fixing algae, but the phosphate had to have a mineral source. Chemical weathering—water washing over phosphate-rich rocks— dissolves and supplies part of the phosphate in aquatic ecosystems. The erosion of small soil particles supplied another part. And as the decomposing algae yielded carbon dioxide, nature’s recycling system caused the ooze to yield phosphate back into the water, providing nutrients for another generation of vegetation. The dead and decomposed detritus mixed with fine clays and became the ancestor of oil shale. The discovery of fossilized bottom-feeding catfish in Lake Gosiute oil shales indicates there were dynamic alternating periods of aerobic and anaerobic conditions on the bottom of the lake and that the deposition of oil shale did not depend on a permanent lake-bottom environment without oxygen. The algae blooms became the primary producers for a rich organic ecosystem. Free-swimming algae (phytoplankton) fed the filter feeders. Fixed algae (periphyton) fed the browsing and grazing community at the lake bottom (benthos). Larger submerged, floating, and emergent vegetation contributed their share of carbon. The decomposers did their work at the lake bottom, while the carnivores fed on the filter feeders, the grazers, and the browsing herbivores. Part of the carbon and part of the nutrients were recycled, but the net result of this complex biological-biochemical process was to fix CO2 from the atmosphere and incorporate it into lake sediments as organic carbon. As the lake dried up, it buried the organic ooze. That ooze was extraordinary for several reasons. First, it was remarkable for the rate and area of accumulation. Plant growth must have been astonishing. Two factors helped to boost the plant activity. Atmospheric CO2 concentrations were significantly higher and Eocene lake temperatures were significantly warmer than today—perhaps as much as 10°C—and the lakes were always above freezing.8 Second, the ooze accumulated in vast numbers of layers over the lake’s long history. The inter-

mingled layers of organic-rich (kerogen) and organic-poor sediments (sometimes carbonate-rich and sometimes saline) that accumulated at the bottom of Lake Gosiute solidified over eons and are now called the Eocene Green River Formation. Plant materials, dead fish, diatoms, dust, sands, silts, and clays rained down from the overlying waters to the bottom of the lake and enriched the layers. Limnologists use these layers as annual indicators of climate, environmental change, and lake productivity. The summer/winter pairs of sediment layers are called varves. At Elk Lake, Minnesota, varves have marked off the seasons annually for the last 10,400 years. These varved sediments yield abundant biological, chemical, and geological information that can be related to climate and environmental change. Varve analysis allowed limnologists (scientists who study freshwaters) to discover that the mixed hardwood forest surrounding Elk Lake was first a spruce forest, then a pine forest that began to be replaced by prairie vegetation dominated by grasses, sagebrush, and oak as the area became markedly drier.9 Similarly, Lake Gosiute varves, now exposed layers of sandstone, siltstone, and shales of the Green River Formation, are amazing storehouses of information. Volcanic sands from the Absaroka volcano field region near Yellowstone are found in the Red Desert and mark a period when Lake Gosiute, like Vermillion Creek, stole water from the Wind River basin. Water and sediment were transported over the lower divides, and titanic volumes of water again filled Lake Gosiute, perhaps to its greatest extent. Again aquatic vegetation flourished and great quantities of carbon dioxide were fixed by photosynthesis, leading to the deposition of precursors of kerogen-rich oil shale. In the nutrient-rich backwater bays of warm Lake Gosiute, photosynthesis proceeded at warp speed. It created a carbon reserve— one solution to a carbon energy crisis created by a biped species that would walk the planet in the future. Beach sands in the eastern part of the then-freshwater lake were left exposed as the lake retreated eastward and stood still. On the abandoned shoreline sands the peat bog grew quickly, punctuated by marshes

Water in the Red

109

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 110

a few feet higher. Yet another westward expansion of the lake inundated the bog, and the sunken stage was set for the formation of the Wamsutter Rim canneloid coal bed. Twelve square miles of bogs were flooded and became the fodder for anaerobic decomposition that metamorphosed into a five-foot-thick, four-mile-long coal deposit, amounting to some 68 million short tons. A westward expansion of the lake washed over beaches, trending on a north-south axis, and transformed the earlier smaller lake in the southern part of the Red Desert into a very large lake that occupied nearly all of the present Green River basin. The outlet of the lake then closed, and once again the lake waters became very saline. The Great Salt Lake in Utah is as closely analogous as any lake we see today in North America.10 Oil Shale Lake Gosiute led to the deposition of the world’s largest known deposit of oil shale. Oil shale is a mixture of fine clays, alkaline carbonates, and organic materials that are derived from plant materials photosynthesized from a rich CO2 atmosphere. In Wyoming these shales were deposited where still lake conditions allowed the clays to settle along with the organic matter that would decay and become kerogen—the fuel of oil shale. The Green River oil shales can yield 65 to 70 percent of their organic matter as oil.11 Even though it played a producer’s part in the creation of a world-class deposit of oil shale, Lake Gosiute was not done. The oil shale is closely associated with another world-class deposition—the trona beds.12 Trona Much happened in the 4 million years Lake Gosiute lived, draining some 48,550 square miles. Imagine standing on the shore of shallow Lake Gosiute and looking toward the summer setting sun. The lake is rimmed with immense mudflats. Behind you is the rocky outcropping of the Rock Springs Uplift. The lake stretches miles to the horizon as you pan from north to south. To 110

the southwest, broad, gently sloping alluvial plains cascade toward the lake, separating it from the mountains. The slopes are punctuated by ponds and swamps. The clay mudflats at your feet are red—the red beds of the Red Desert.13 There are palm trees that will die as the water becomes saltier; they will fall and, once buried in the mud, will flatten along their length. Fossilization will come much later as warm groundwater saturated with silica invades and replaces the carbon cell structure. Once fossilized, the palms will be buried and resurrected hundreds of times by wind and by smaller meandering creeks for millions of years. As you stand watching the sun set over Lake Gosiute, it is warm and windy but not tropically humid. You are closer to the equator, but more importantly, the elevation is low — perhaps 1,000 feet. The mean annual temperature is 19°C, or 67°F.14 The Lake Gosiute climate was much warmer than today’s, and the area could have experienced temperatures far higher than any modern world record. Little wonder that the evaporation rate played such a critical role in the lake’s history. Early estimates of the precipitation near the lake were about three feet per year, but evaporation was almost four feet annually.15 Later, the temperature and precipitation ranges for the Eocene were reinterpreted, based on vegetation pollen. The early estimates square well with the reinterpretation, but as the Eocene progressed, the climate cooled a few degrees and the precipitation decreased markedly. Seasonal changes were much more pronounced but, in the subtropical climate, that did not mean freezing. The area remained humid enough to support luxuriant vegetation in freshwater areas.16 Flash forward through time as the temperature increases, precipitation decreases, evaporation increases, and Lake Gosiute retreats away from your feet and toward its center, an almond-shaped concave depression. As the lake recedes, it increases in salinity. Salts effloresce as the mudflats dry out. Rare storms wash sheets of water over the flats, and the salts dissolve and are carried to the depression. Finally, the salinity in the remnant part of old Lake Gosiute reaches the tipping point that

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 111

chemists call saturation, with carbonates of sodium coprecipitating as the mineral trona. Augmented by salts found in local springs and salts dissolved from volcanic ash, the rivers carried enormous amounts of the substances into the lake.17 As the mud dried, the dissolved salts crystallized on the circumference in the process of efflorescence. The salty brine, generated on the fringing mudflats by episodic wetting and drying as the lake expanded and shrank, was concentrated by something hydrologists call evaporative pumping. If you were to saturate a soil with brine and then allow it to evaporate, capillary action and osmosis would move not only the water, but also the salts dissolved in the water, to the surface. These crystals must have given the lake a frosty halo appearance. Seasonal flooding washed and dissolved the salts into the lake’s center. Climatic change drove Lake Gosiute’s fluctuation in size.18 At its lowest level Lake Gosiute was possibly never less than 3,800 square miles. However, the bedded salts formed only in a much smaller area in the southern part of the Red Desert.19 The retreat toward the lowest point in the lake likely left areas of Gosiute isolated and less saline. Eventually, the lake fell so low that the southern part was completely cut off.20 The evaporation must have continued in this isolated portion of Lake Gosiute. Inexorably, the brine became more and more concentrated, periodically receiving salty recharges when Lake Gosiute levels rose. Those recharges depleted the salts from the more remote parts of the lake. Perhaps isolated bays in the northern and eastern part of the lake were connected to the lowest part of Lake Gosiute, and saline water continued to flow to the southern part of the lake. During wetter periods the lake likely rose in the northern and eastern arms and dissolved and carried additional salts into the lower parts of the lake, where they accumulated. Drier periods caused the accumulated salts to concentrate by evaporation and, likely in some areas, to precipitate. Given 4 million years to repeat this pattern of evaporating, isolating, dissolving, and transporting salts, Lake Gosiute produced a reserve of saline minerals that is unique in the world.

The reserves were deposited in the lower isolated southern portion of the lake, now known as the depocenter. Perhaps as much as 100 billion tons of the stuff was deposited in as many as forty-two distinct beds over an area of 1,300 square miles.21 Eighteen principle beds in the southwestern Red Desert, ranging in thickness from about 3 feet to almost 38 feet were left behind as Lake Gosiute exhaled its humid breath into the hot, arid desert. The materials weren’t deposited only in beds. Disseminated and incorporated through the muds of Lake Gosiute, augmented by CO2 from biological activity, and concentrated by evaporation, the mineral trona occurs as individual crystals and larger crystal clusters alongside the saline minerals. The salts had three sources: salinity dissolved and carried to the lake in rivers; mineralized springs; and salts leached from ashfall springs.22 But these were not ordinary salts. The salts left behind were carbonates. The carbonate salts of sodium come in two flavors: sodium carbonate, commonly called soda ash, and sodium bicarbonate, or baking soda. Deposited in the evaporating Lake Gosiute, the carbonate salts occur in equal abundance along with two molecules of water— called waters of crystallization, or hydrates. The soda ash, baking soda, water, and impurities make up the mineral trona. Simply heating this mineral causes it to decompose into soda ash. The water driven off by the decomposition most certainly is water from Lake Gosiute held captive in the mineral for 50 million years. Driving past the trona refineries on Interstate 80 in southwestern Wyoming, few passersby are aware that part of the steam they see is water from Lake Gosiute — liberated and moving again in the hydrologic cycle. How much of Lake Gosiute was fixed in the deposition of trona? For 100 billion tons of trona, 16 billion tons of Lake Gosiute was fixed for geological time. Sixteen billion tons of water is enough to fill Flaming Gorge Reservoir three times. Four factors drove the deposition of trona: temperature and the concentrations of chloride, sodium, and dissolved CO2.23 During the most intense periods of trona deposition, Lake Gosiute did not evaporate to dryness; otherwise sodium chloride, much more

Water in the Red

111

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 112

soluble than sodium carbonate, would also have formed. The trona beds are noted for their purity. Most importantly, if Lake Gosiute deposited 100 billion tons of trona, the deposition fixed 38 billion tons of CO2 from the atmosphere. Along with the tremendous quantity of CO2 fixed by photosynthesis and locked up as oil shale and coal in the Green River Formation, the CO2 fixed as the mineral trona certainly contributed to an anti-greenhouse effect. Globally, the fixing of CO2 by vegetation and similar mineral processes (called lithification) is why the climate is much cooler today than in Eocene times when Lake Gosiute was alive. 112

The eastern bay of the lake was probably always less saline because of two large rivers that emptied into it. One, at the northern end, drained the Granite Mountains to the northeast of the Red Desert. The other drained a large part of the eastern Uinta Range, emptied into the lake from the south, and presumably left deposits like Pine Mountain. Remarkably few, if any, salt deposits have been found in these areas.24 The jagged crescent-shaped Rock Springs Uplift became an island in Lake Gosiute. As the lake evaporated and lost volume, the island became a peninsula. With further shrinking, the uplift became the eastern bank, then

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 113

a vantage point from which one could have seen the retreating salt-encrusted shoreline turn red in the light of the setting sun. Much can happen in 4 million years. As the mountain ranges forming the rim of the hydrographic divide rose, the sedimentary floor of the basin fell, and with it, Lake Gosiute. The lake dried up for the final time, perhaps as a result in part of the fixing of CO2 and the cooler and drier climate. Vegetation that once slowed erosion probably died off in the cooler, drier climate. Erosive forces then deposited hundreds to thousands of feet of sediments on top of the layered salts and shales. Eventually an outlet was cut through the southern fringe mountains, and relatively pip-squeak rivers began moving the Red Desert segments again—this time out of the basin. Black Trona Water In 1992 a well driller contracted to the Bureau of Reclamation went about his work on the banks of a stream in the arid northwest corner of the Red Desert, a place aptly called Bone Draw. Bone Draw is a stressed and crooked arm of the Big Sandy River, a tributary to the Green River–Colorado system. The well driller’s task was to find a new freshwater source to augment and dilute the irrigation return flow in Bone Draw. He found instead a root beer– colored geyser of alkaline water that drenched his rig, quickly eroded and enlarged the bore, and ran uncontrolled for weeks. The appearance of the black water was “accompanied by a small amount of a combustible gas, leading the driller to suspect flowing oil,” as observers reported. They added, “This illusion quickly dissipates when he touches the soapy-feeling liquid, particularly with a cut finger, whose stinging promptly reports that this is not oil.”25 The driller was searching for a freshwater aquifer to assist a local conservation group—Trout Unlimited—in its efforts to turn Bone Draw and the Big Sandy River into trout-rearing streams. Locals maintained that if the Bureau of Land Management (BLM) had only asked their advice, they could have warned the bureau of these “blowout” wells of what they called black trona water.

Bureau of Reclamation hydrologists were mystified. No other known aquifer in the world had such water. Where was the water coming from? What turned the water into a fountain of venting pressure? What gave the water its unique and bizarre color and caustic chemical character? This was indeed a bit of a mystery. In 1938 the first discovery of this curious brine had been made on the John Hay ranch south of Eden, Wyoming. This well also flowed out of control for weeks and foamed “like brown soap suds for hundreds of yards down the draw.”26 While exploring the extent and quality of trona deposits in the Eden/Farson area, John McDermot, of the Wyoming Trona Company, used a cable tool machine to drill a well hole. Black water flowed from that hole at an estimated rate of 400,000 gallons per day.27 At least twenty-six wells have produced the black water. All have been drilled in the high, cold valley of the greater Red Desert called Eden. Nearby wells drilled to the same depth (400 – 600 feet) did not always yield black trona water, even though wells drilled tens of miles apart to variable depths produced copious flows of the mysterious fluid. Thousands of wells have been drilled in the Red Desert, and probably a thousand exploratory wells were drilled in the trona beds alone. Why did the black trona water occur only in these isolated places? A single glance anywhere in today’s Red Desert convinces the viewer that any water is scarce and an artesian water source is most rare. Water from the Big Sandy River originates from the melting snows and rainstorms in the southern Wind River Range. Big Sandy water irrigates the only farming region in the Red Desert — the Eden-Farson project. The project brought a new reservoir and modern-day homesteaders who won a lottery following World War II. Eventually 20,000 acres of irrigated land came into agricultural production in this region. Today’s Red Desert has more than location in common with the Lake Gosiute valley. Salts, the same salts that Lake Gosiute left as its legacy, still flavor the soils and waters of the Red Desert. Salts that are mined in the “Trona Patch” and bring Red Desert

Water in the Red

113

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 114

prosperity also cause problems when they accumulate in irrigation return flows. Salts that deposited in great thick lenses, now hundreds of feet underground, are present in lower concentration on the surface. Irrigation water dissolves them. To prevent salt accumulation, more irrigation water is applied in the technique of flood irrigation. Irrigation return water flushes these salts into the Big Sandy through drainages like Bone Draw. Salinity is an interstate concern along the Colorado River, and dedicated government efforts to reduce the salt load have encouraged farmers to switch their irrigation method to pivot sprinklers. But the cola-colored, slick-feeling artesian water was not what trout could use for rearing their young—indeed, it has a salinity greater than that of ocean water. And the color? That comes from the dissolution of organic acids like humic and fulvic acids.28 The black trona water found in artesian abundance has no current practical uses, although one well was reported to flow at 276,000 gallons per day and could yield 62 tons of soda ash and 130 tons of organic acids each day. The chemical nature of black trona water hints at its origin. It must have been in place for eons and dissolved not only trona but many organic compounds as well. The black trona aquifers lie underneath an impermeable oil shale layer, so the source of the water could not be surface water infiltrating from above through fissures and pores. Further, if the aquifer occurred in the trona district to the southwest, it would lie just above the highest bed of trona. In addition, the water is artesian and pressurized by the weight of the earth above it; geologists call this “overburden pressure.” Scientists suggest that the brine may have formed in the depocenter of old Lake Gosiute and, confined by the oil shale layer above it, migrated up the slope toward the mountains. Water confined in a salty sedimentary layer through time will certainly dissolve all the salt that geochemical variables allow. Dissolved trona accounts for the caustic character of the water. Curiously, the organic compounds that dissolved in the alkaline trona water are not soluble in organic solvents, and yet, when heated, about three-quarters of the mass turns into oil. Could the black organic compounds have been leached from the 114

oil shale? They could not, for oil shale will not yield organic acids when they are extracted with alkaline water containing trona.29 If the water cannot infiltrate from above, could it come from below? The black trona water aquifer is confined by oil shale, which has virtually no permeability—suggesting that the water was there when the shale was formed.30 What then was the source of the water, and how did it get trapped in the oil shale? One clue lies in the organic acids that can form in lake sediments in the presence of small amounts of oxygen. These organic acids could have formed in the interstitial spaces, the tiny water-filled and organic-rich voids in the sediments of Lake Gosiute. A depression in the lake bed could harbor just the right conditions to produce the organic acids, store the highly alkaline (dissolved trona) water until something, perhaps a volcanic ash, sealed the aquifer. Sealing the aquifer would be required to preserve the water and keep it from being expelled as more sediments accumulated on the lake bottom and the pressure of their weight squeezed the water out of the sediments. Sealing the aquifer would also be critical to protect the organic compounds from the reducing environment that produced the kerogen in the oil shale. All of the Lake Gosiute sediment water was probably not “preserved” at the same time; rather, its preservation occurred, much like the precipitation of the trona, in a long series of recessions and advances of the lake.31 And so some water of Eocene Lake Gosiute still exists today, trapped in trona crystals and in an ancient aquifer mere miles from its origin, while the North American continent itself has moved hundreds of miles. The steam seen rising from the soda ash refineries liberates Lake Gosiute water back into an atmosphere much different from the atmosphere from which it precipitated 46 to 50 million years ago. The black trona water remains buried in its dark grave, ready to spring to the surface and yield its secrets. Water in the Red Today All continents except Antarctica have a continental divide. North America’s continental divide is the quintessential example of a

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 115

hydrologic divide—a physical boundary defined by elevation and topography—upon which, if one tossed a bucket of water evenly, half would flow into one basin and half would flow into the other. In North America the divide separates waters that flow into the Atlantic Ocean via the Gulf of Mexico from those that flow into the Pacific Ocean. As the continental divide moves northwestward from Colorado’s Routt National Forest toward the Red Desert, it looks like a computer-generated line in fat-bits mode. Falling in elevation, the continental divide follows the crest of a series of ridges separating the North Platte River drainage from the Little Snake River drainage. After entering the Red Desert proper, the divide splits in eastern Carbon County and circumnavigates the Great Divide basin, a unique internally drained feature of the Red Desert with saline lakes, ponds, and large playas that alternate between mudflats during wet years and salt pans in droughts. Constituting 4 percent of Wyoming’s surface area, the Great Divide basin is large—2.5 million acres—and is the only place in the United States where the continental divide splits to form such a large endorheic watershed. An endorheic watershed is one in which there is no flow of water from inside the basin to the outside either on the surface or through groundwater interflow. Any precipitation falling inside an endorheic watershed reenters the hydrologic cycle only by evaporation, which is typically high in this windswept region. The captive water not lost to evaporation in the Great Divide basin has a long residence time and, consequently, a long time to dissolve salts both on the surface and underground. The closed nature of the Great Divide basin hydrology means that salts and minerals leached from rocks in the surrounding terrain also stay in the basin and are concentrated when the solution water evaporates. Lake Gosiute for most of its life resided in an endorheic basin. The same processes that created Lake Gosiute and the trona beds are at work today in the Great Divide basin. There are no permanent streams in the basin—just intermittent arroyos that track historical snowmelt and thundershowers. Water flows to the lowest point, with an elevation around 6,500 feet. Here we find

several gossamer water features, made fleeting by the vagaries of the balance between precipitation and evaporation. Red Lake, Hay Reservoir, and the Chain of Lakes are not permanent lakes and reservoirs. They change in size and shape dramatically from year to year and shrink markedly, often breaking into several smaller bogs wetted with less than a foot of runoff —hence the name “Chain of Lakes.” This wetland complex of capillary springs, playa lakes, and ill-defined clay mud bogs has mired many a traveler. What appears to be a dry, unvegetated flat is a desiccated crust that cannot support weight and transforms into a grunge-gray, sticky, sucking dense clay mass that will trap a vehicle in seconds. So stuck, the only hope is to get a tow from an infrequent passerby. To many this broad, relatively treeless landscape of the Great Divide basin is the epitome of the Wild West. To avoid the quagmire of saturated surface soils, the California-Oregon-Mormon Trail passed to the north of the Great Divide basin, and the first transcontinental railroad routed to the south. However, as development has expanded into the basin, the ephemeral stream network has been disturbed by the construction of dams and roads. Building roads causes linear traps of rare precipitation flowing overland. Millennia of creeping overland water flows have had just enough energy to loft the small clay particles and fill the basin low spots. These clay flats become the bait for unsuspecting victims anxious to test their new off-road vehicles. Traffic closes the pore spaces in road materials and summer thunderstorms wet the clay into a slick coating that sometimes breaks loose from drier material and is thrown up between tire and fender, packing in as solid as concrete. Where the roads cross drainages, a standardized culvert as required by the BLM or other agency often either gets clogged with soil and debris from “gully washers” or becomes the source of downcutting of the stream channels. This downcutting is caused by the increase in stream velocity within the culvert and the consequent increase in the erosive force of the stream. Without any armor on the banks of the receiving channel, the now more en-

Water in the Red

115

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 116

ergetic stream starts cutting at the banks and bed of the stream. The streams in the Great Divide basin are also all ephemeral, flowing only in response to snow melting on frozen or watersaturated ground, or in response to summer thunderstorms that overload the land’s capacity to absorb the falling water.32 Often, falling raindrops hit the ground with such force that small craters form, and the ejecta—small clay particles appearing as powder— from the jillion craters can be seen as a curious dust rising just as the storm begins. The distribution and extent of the summer thunderstorms are usually limited to a few square miles or less. The storms appear as gray brush marks under giant clouds that transition from brilliant white popcorn balls on top to black horizontal schisms near the basin floor. In the early 1980s, standing on the rim of Steamboat Mountain and overlooking the Great Divide basin, Luna Leopold—University of California professor emeritus of geology and the world authority on fluvial geomorphology — remarked that with the single exception of the volcanic rocks underneath, all of the landforms we could see had been shaped by water and wind alone. “Streams do their work at bank-full stage,” he said, “and the rugged landscape we see is not the result of catastrophic floods and calamitous windstorms, but the result of countless small events recurring over millions of years.”33 As the runoff and the suspended and transported sediment move downstream away from the storm, the channel absorbs more of a stream’s energy and the dry streambeds absorb more of the water. This decreases both the velocity and volume of the stream and dramatically reduces its energy. Exhausted and no longer able to transport the sediment it has picked up, the stream deposits it along the bed and occasionally on top of the banks. It is not unusual to see flows of several hundred cubic feet per second at speeds faster than a human can run, coursing down arroyos carrying water that looks like café au lait. In this manner the alluvial flats are built and basins filled over geologic time. Mimicking pools and riffles in rivers, the sediments in ephemeral channels are deposited as gravel bars in nearly identi116

cal frequency of spacing to that of pools and riffles—five to seven channel widths along the length of the stream.34 The same general kinetic processes that cause riffles, pools, and bars in perennial streams can be said to account for this pattern in ephemeral streams. These streams also move larger rocks. As water spills into the channel from overland, the energy is the product of the volume (actually the mass) and the velocity squared. Thus, water that is moving twice as fast has four times the erosive force. This force picks up the lightest clays first and then the larger material. Gravel that it cannot suspend is bounced along the bottom. When the water encounters a rock so large it cannot be moved, it flows around it. Like the air moving around a wing in flight, the water moves faster around the rock than in the channel. This increase in kinetic energy causes localized erosion around the large rock and the rock “falls” in the direction of the stream flow. The bed is thus scoured in places and filled in others. As this sequence is repeated incalculable times, the ephemeral channels in the Great Divide basin wander about and build the basin floor in processes identical in physics to those that have carved and filled the Red Desert. Hydrologists characterize watersheds according to their size. Within the greater Red Desert there are five watersheds in addition to the Great Divide basin.35 The upper Green River is the largest of the five—the great drain marking the western boundary of the Red Desert. The Green River receives the flow from all Red Desert runoff outside the Great Divide basin. After being discharged from temporary captivity in Flaming Gorge Reservoir, the Green welcomes the confluence of Vermillion Creek in Colorado and Muddy Creek in Utah via the Little Snake River and the Yampa River. In Wyoming the Green receives the waters of both the Big Sandy River—which is roughly the northern boundary of the Red Desert—and Bitter Creek. Snowmelt is the source of most runoff in the Red Desert. As in other arid western watersheds, irrigation is the largest consumptive use of water in the Red Desert, and almost all irrigation uses surface water. Because most of the streams are ephemeral, few have been studied extensively. Two notable exceptions are

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 117

Salt Wells Creek, a tributary of Bitter Creek that originates on Pine Mountain, and Separation Creek in the extreme southeastern corner of the Great Divide basin.36 Streams that drain the higher-elevation areas of the Red Desert usually peak in May, but rapid early snowmelt, often accompanied by warm spring rain, caused substantial flooding in Rock Springs in 1924, and summer rains flooded Rock Springs again in 1937.37 Surface water quality is good at the higher elevations but decreases rapidly downhill as the water dissolves salts and picks up suspended sediment in the streams’ flatter reaches. The names of Bitter Creek tributaries (Killpecker Creek, Salt Wells, Alkali Springs) give away pioneers’ opinions of their quality. Sediments eroded from parent soils in the Bitter Creek drainage are richer in nutrients such as phosphorus than the mean soils themselves.38 This means not only that the erosion is removing the most nutrient-rich soils from the drainage, but also that the eroded soils cause water quality problems in Flaming Gorge. The phosphorus causes an accelerated aging, called eutrophication, in the reservoir. Eutrophication is a naturally occurring process in lentic water bodies, normally very slow but often accelerated by human activities. Anything that increases erosion in the Red Desert like overgrazing, road building, culvert installation, and oil and gas production will ultimately end up as a problem in Flaming Gorge. As a consequence, in warm-water years the riverine portions of Flaming Gorge Reservoir often look like pea soup as algae populations grow exponentially. Immerse your arm in water at the end of the Firehole Canyon boat ramp in mid-September, and most years you cannot not see your upturned hand because of the water turbidity. Excessive nutrient concentrations stimulate the growth of undesirable algae. Other plants, such as water weeds and slime organisms, contribute to the nuisance growth. Given the perfect conditions of high water temperatures, high nutrient concentrations, and all the water they need, these plants, through the miracle of photosynthesis, consume large amounts of carbon dioxide during the day and give off oxygen. When the respiration cycle

reverses itself during the night, dissolved oxygen levels sink precipitously and desirable fish species like trout are seriously affected. Clearly, eutrophication in Flaming Gorge Reservoir, named by John Wesley Powell for the same red-banded coloration of the Red Desert, has caused the reservoir to become a fishery in transition.39 Similar processes on a grand scale began the fixing of carbon in Lake Gosiute 50 million years earlier, the difference being that the water was then captive in the Red Desert basin. Bitter Creek is one of the few drainages that has no mountain source, originating instead in the higher elevations (6,700 feet) east of Fort LaClede and on the north side of the Haystacks — a remote area of the southern Red Desert. Bitter Creek has the distinction of being one of the largest watersheds of an ephemeral stream in the United States—a surprising 2,220 square miles.40 By comparison, the Gordon River of Tasmania has a watershed of almost identical size and a perennial average flow of 10,300 cubic feet per second, whereas Bitter Creek dries up completely each year.41 While groundwater in the Red Desert is present almost anywhere below a certain depth, the quantity is limited and the amount of total dissolved solids— or mineral salts—in the water almost invariably exceeds the U.S. Public Health Service standard of 500 milligrams per liter, often by a factor of three or higher. Because of Wyoming’s participation in efforts to limit the discharge of saline water to help comply with the Colorado River Basin Salinity Control Act, discharge of produced saline water into surface drainages may ultimately prove to be a limiting factor in the development of gas resources, particularly coal-bed methane development. Removal of dissolved salts from water is an expensive and vexing problem both in Wyoming and internationally. As a consequence of open-pit uranium mining, years of leaking tailings ponds have polluted groundwater in the northeastern Red Desert. This has proved to be particularly difficult to reverse, and companies have petitioned to be allowed to complete reclamation without restoring contaminated groundwater.42

Water in the Red

117

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 118

Notes 1. R. T. Leigh, Wyoming Trona: An Overview of the Geology, Proceedings of the First International Soda Ash Conference, vol. 2, Wyoming State Geological Survey Public Information Circular 40 (1998). 2. Dennis H. Knight, Mountains and Plains—The Ecology of Wyoming Landscapes (Yale University Press, 1994), 12. 3. W. H. Bradley, “Geology of the Green River Formation and Associated Eocene Rocks in Southwestern Wyoming and Adjacent Parts of Colorado and Utah,” U.S. Geological Survey (hereafter “usgs”) Professional Paper 496-A (1964), 88. 4. W. H. Bradley, “Origin and Microfossils of the Oil Shale of the Green River Formation of Colorado and Utah,” usgs Professional Paper 186 (1931), 58. 5. R. C. Surdam and C. A. Wolfbauer, “Green River Formation, Wyoming: A Playa Lake Complex,” Geological Society of America Bulletin 86 (1975): 335–345. 6. H. P. Eugster and R. C. Surdam, “Depositional Environment of the Green River Formation of Wyoming: A Preliminary Report,” Geological Society of America Bulletin 84 (1973): 1115–1120. 7. D. E. Garrett, Natural Soda Ash (Van Nostrand Reinhold, 1992). 8. Matthew Huber and Rodrige Caballero, “Evidence for Robust Tropical Dynamics in the Hot House,” Science 299, no. 5608 (2003). 9. W. E. Dean, “Evidence of Climate Change over the Last 10,000 Years from the Sediments of Lakes in the Upper Mississippi Basin,” usgs Fact Sheet FS-059-99 (April 1999). 10. W. C. Cuthberson, “Stratigraphy of the Trona Deposits in the Green River Formation,” Contributions to Geology, University of Wyoming, Laramie (Spring 1971). 11. H. N. Smith, J. W. Smith, and W. C. Kommes, Petrographic Examination and Chemical Analyses of Several Foreign Oil Shales, U.S. Bureau of Mines Report Inv. 5504 (1959), 1–34. 12. R. C. Surdam and C. A. Wolfbauer, “Origin of Oil Shale in the Green River Formation,” in Wyoming Geological Association Guidebook, Twenty-fifth Field Conference (1973), 207. 13. Ron C. Surdam and K. O. Stanley, “Lacustrine Sedimentation during the Culminating Phase of Eocene Lake Gosiute, Wyoming (Green River Formation),” Geological Society of America Bulletin 90 (1979): 106. 14. W. H. Bradley, “The Varves and Climate of the Green River Epoch,” usgs Professional Paper 158-E (1929), 92 –93. 15. Ibid., 93. 16. Henry R. Roehler, “Eocene Climates, Depositional Environments, 118

and Geography, Greater Green River Basin, Wyoming, Utah, and Colorado,” usgs Professional Paper 1506-F (1993), 2. 17. Surdam and Stanley, “Lacustrine Sedimentation,” 93–110. 18. W. H. Bradley and H. P. Eugster, “Geochemistry and Paleolimnology of the Trona Deposits and Associated Authigenic Minerals of the Green River Formation of Wyoming,” usgs Professional Paper 496-B (1969). 19. Ibid., 22. 20. Ibid., 24. 21. Cuthberson, “Stratigraphy of the Trona Deposits.” 22. Bradley and Eugster, “Geochemistry and Paleolimnology.” 23. Ibid., 47. 24. Ibid., 22. 25. George F. Dana and John W. Smith, “Black Trona Water, Green River Basin,” Wyoming Geological Association Guidebook, Twenty-fifth Field Conference (1973), 154. 26. Leonard Hay, pers. comm. 27. Dana and Smith, “Black Trona Water,” 153. 28. John R. Dyni, Sodium Carbonate Resources of the Green River Formation, usgs Open-File Report 96-729 (1996), 22. 29. Dana and Smith, “Black Trona Water,” 155. 30. G. F. Dana and J. W. Smith, “Nature of Black Trona Water Occurrence, Northern Green River Basin,” Wyoming Geological Association Earth Science Bulletin 9 (1976): 9 –15. 31. Ibid., 12. 32. “Ephemeral” is the word hydrologists use to describe streams. “Intermittent” is used to describe springs that vary markedly in output— the most famous of which is the Intermittent Stream in western Wyoming’s Star Valley. 33. Luna B. Leopold, pers. comm., Wyoming Outdoor Council field trip, June 1982. 34. Luna B. Leopold, M. Gordon Wolman, and John P. Miller, Fluvial Processes in Geomorphology (W. H. Freeman and Co., 1964), 209. 35. P. R. Seaber, F. P. Kapinos, and G. L. Knapp, “Hydrologic Unit Maps,” usgs Water-Supply Paper 2294 (1987). 36. H. W. Lowham, D. A. Peterson, L R. Larson, E. A. Zimmerman, B. H. Ringen, and K. L. Mora, Hydrology of Area 52, Rocky Mountain Coal Province, Wyoming, Colorado, Idaho, and Utah, usgs Water-Resources Investigations, Open-File Report 83-761 (1985). 37. R. T. Rhode, Booms and Busts on Bitter Creek (Pruett, Boulder, Colo., 1987), 139. 38. C. D. Thompson, “Nutrient Removal from Arid Lands Runoff by

Craig Thompson

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 119

Settling,” Forum of the Association Arid Land Studies 4 (1988): 144–151. 39. B. Wengert, “Flaming Gorge—A Fishery in Transition,” Annual Lecture Series for Ways of the River, Western Wyoming Community College, Wyoming Game and Fish Department, June 6, 1996. 40. Thomas Wilkinson, project engineer, Lowham Engineering llc, Lander, Wyoming, pers. comm., June 14, 2007. gis analysis based on Hydrologic Unit Code 12 Watershed Boundary, usgs, 2002. 41. Luna B. Leopold, A View of the River (Harvard University Press, 1994), 107. While this is true above the city of Rock Springs, anthropogenic flow of wastewater below Rock Springs to the confluence of the Green River gives the stream an artificial appearance of perennial flow. 42. J. D. Collins, “Reclamation and Groundwater Restoration in the Uranium Milling Industry: An Assessment of umtrcva Title III,” Journal of Natural Resources and Environmental Law 11, no. 1 (1996): 23.

Water in the Red

119

Proulx_03_pp.073-120:Layout 1

8/4/08

6:17 PM

Page 120

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 121

3. Environmental Change in the Wyoming Basin’s Red Desert Dudley Gardner

ometime in 1870 a soldier planted a tree along officers’ row at Fort Bridger. In 1890 the military stopped watering that tree. Its rings show the drought of the 1930s and even a drier period in the 1950s.1 The tree fell in the early 1990s. Contained in its rings is the record of 100 years of changing environment, showing fluctuations in moisture and aridity, a cycle of wet and dry periods. The soil history of the Red Desert indicates that, long before humans came, the region was becoming increasingly arid. When people did arrive in the area, between 10,000 and 12,000 years ago, sand dunes had formed, creeks had dried up, and soil had settled downriver or eroded in heavy runoff.2 How has the environment changed since people began to call the region home? From 20,000 to 15,000 years ago, the area experienced glaciation from the Wind River and Uinta mountains (see Table 3.1).3 Unlike today when the dominant winds come from the southwest, during the last ice age prevailing winds came from the northeast.4 Summers were much colder, and winters cold and moist. Pine trees, buffalo berries, and serviceberries grew in many areas, and cattails surrounded the numerous small lakes. Grasses were taller, eaten by megafauna like mammoth, large bison, camels, horse (Equus), and large antelope.5 Sagebrush covered great expanses. The primary distinction of the ice age was the

S

abundance of water within the Wyoming Basin. The surrounding mountains still created a rain shadow, but enough moisture reached the basin to make the area much wetter than it is today.6 The end of the ice age began 15,000 years ago with changes in the earth’s orbit and axial tilt. Following Milankovich’s cycles, this led to increased intensity of cold and warm seasons.7 But neither Pacific subtropical weather patterns nor monsoon flows developed in southwestern Wyoming.8 Aridity increased. The region was cooler and drier, probably because of a change in wind patterns or a shift in the jet streams.9 Deserts are created when there is increased evapotranspiration and greater wind speeds over long periods, combined with drier weather patterns. The change in wind and water patterns altered the Wyoming Basin. The Killpecker dune fields started to take shape during the last ice age, and the formation process accelerated at about 12,500 bp (before the present).10 Dune fields in general “are considered to contain records of climate change, with eolian sedimentation and erosion indicating trends toward arid conditions.”11 The arrival of Native Americans in the basin by around 12,000 bp corresponded with an increase in moisture.12 From 11,000 to 9000 bp, the wettest period in the postglacial era swept the region. Around 11,000 years ago the jet stream shifted northward, producing “stronger monsoonal flows from the Gulf of Mexico 121

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 122

Table 3.1. Generalized climate regions from 20,000 to 150 BP Years before presenta

Miller modelb (years before present)

20,000 –15,000

Description

Full glaciation—Laurentide Ice Sheet

15,000 –11,000

13,000 –11,000

Deglaciation— cooler; severe aridity

11,000 –10,500

11,000 –9,500

Most mesic interval of postglacial era

10,500 –9,000

9,500 – 6,500

Increased aridity (early Holocene)

9,000 – 4,000

6,500 – 3,500

Eckerle sees this as a very dry period peaking between 7,500 –5,000 (middle Holocene). During this time there were episodes that exhibited more moisture, but in general this was a drying period. Miller sees this as a period with more cyclical variability in both dry and wet periods.c

4,000 – 3,500

3,500 –2,800

Similar to present

3,500 –1,800

2,800 –1,800

Neoglacial (moist)

1,800 –1,100

1,800 –900

Similar to present

1,100 –900

Moister

900 –500

900 –500

Very dry

500 –150

500 –150

Little Ice Age—moist

Source: After Pastor 2003. a. David E. Johnson and Jana Pastor, eds., The Blue Point Site: Paleoindian/Archaic Transition in Southwestern Wyoming (Western Archaeological Services, Rock Springs, Wyo., 2003), Appendix A, 7–18. b. James Miller, “Late Pleistocene and Holocene Geology of the Central Green River Basin with Reference to Geoarchaeology and ‘Benchmark’ Site Selection,” manuscript on file at Kemmerer Resource Area of the Bureau of Land Management, May 17, 1996. c. Jana Pastor, Kevin W. Thompson, Richard K. Talbot, William P. Eckerle, and Erik E. Ingbar, “Seeds-Kee-Dee”: Riverine Adaptation in Southwest Wyoming, Museum of Peoples and Cultures Technical Series, no. 99-3 (Brigham Young University, Provo, Utah, 2000), 419. Miller, “Late Pleistocene and Holocene Geology,” 12 –16.

and the Gulf of Cortez.”13 The changes seem to have created in the interior Pacific northwest, and possibly elsewhere, conditions dominated by wet winters and dry summers. Conversely, in the American southwest, the southern Great Basin, and the Great Plains, moist conditions prevailed. Southwestern Wyoming felt these changes.14 While sand dunes stabilized during this period, the winds that whipped the area still caused mechanical weathering of rock; thus some dunes did increase in size, but not by much.15 The end of this wet period marked the emergence of the drying trend that dominates the region today. 122

At the beginning of the early Holocene (ca. 9000 bp), a drying period began that peaked at about 7500 to 5000 bp. Around 9000 bp, seasonal intensities became pronounced and the area’s summer temperature was 2°C to 4°C higher than today. The water level in the Great Salt Lake fell to its lowest point between 6000 and 3800 bp.16 In general, southwest Wyoming was dry, and xeric plants dominated the landscape. In this period, sand dunes became more active. In the Ferris dune field, 15 meters of sand accumulated at one locality.17 The interval between 7545 and 7035 bp appears to have been the most arid, with twelve to fifteen meters

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 123

of middle Holocene sand evident in the Ferris field.18 While the jury is still out as to what the climate looked like during this period, the steppe environment of sage and xeric grasses that today dominate the region would have been there. From 4000 to 3500 bp there was an environment of alternating wet and dry cycles. One study shows that in the mountains of western Colorado, the upper timberline dropped 100 to 200 m (328 – 656 ft). At the same time subalpine fir dominated the upper elevations, suggesting a shift to drier conditions and winter-dominated precipitation.19 At about 3500 to 1800 bp, a moister and cooler period marked a neoglacial era.20 The plant communities evident today in the Wyoming Basin were likely there at the end of the last ice age, but in this neoglacial period they responded to greater moisture in the soils. There is some disagreement as to what the climate would have been like then, though one intriguing piece of evidence indicates that lake levels in the area reached their highest point between 3,500 and 1,800 years ago. The soils above eolian deposits show there was more moisture. In addition the pollen record keynotes evidence that previously dry areas became wetter in summer.21 Around 1800 to 1100 bp, there was a return to a climatic regime very similar to today’s conditions. Then, around 1100 bp, a warm, wet period encouraged an expanding agriculture. Kenneth Peterson, who looked at the climate of southwestern Colorado, highlights a complex cycle of wet and dry periods at this time.22 He notes that from 1150 bp to 1850 bp and from 850 bp to 625 bp the area was dry and characterized by low winter jetstream precipitation. But from 1375 bp to 1150 bp, and from 950 bp to 850 bp, high winter precipitation marked the southwest. The most intriguing part of Peterson’s model signifies increased summer precipitation from 1200 bp to 1050 bp and a decrease in winter moisture.23 This increased summer moisture correlates to several radiocarbon assays taken from corncobs in granaries in the Douglas Creek drainage in northwest Colorado.24 These radiocarbon assays span the full spectrum of this period, but there may be a jump in radiocarbon assays associated with the years

between 1200 bp and 1050 bp.25 Peterson goes on to note that from 1050 bp to 850 bp there was a decrease in summer precipitation and an increase in winter precipitation. From 950 bp to 850 bp there was an increase in both summer and winter precipitation. Then, from 850 bp to 625 bp there was relatively low precipitation in both summer and winter.26 At 900 bp (ad 1100), an arid period set in that forced people to abandon agriculture in many parts of the west. The period from 900 bp to 500 bp is noted for being decidedly drier. The Salt Lake levels dropped. On the northern plains, eolian activity increased. Xeric species dominated, and the small-mammal collection in Yellowstone suggests a much drier period.27 This dry period endured until 500 bp when the Little Ice Age ushered in a moister, colder period. The so-called Little Ice Age that began 500 years ago persisted until 150 years bp. In general the Little Ice Age was marked by increased advances in alpine glaciers such as those found in the Wind River Mountains.28 Mitchell Plummer contends that moraines from the Little Ice Age are evident from glaciers in the Wind Rivers dating to 1850.29 David Naftz, of the U.S. Geological Survey, found preservation of ice from the Little Ice Age in Upper Fremont Glacier.30 In general, for a glacier to hold its position, there must be more snowfall than snowmelt. This usually occurs when there is more moisture, but specifically it takes place when there are cooler summers and longer winters that increase the potential for more snowfall. Cores from the Wind River glaciers seem to indicate cooler summer temperatures during the Little Ice Age. Tree ring data collected two kilometers from the Upper Fremont Glacier indicate slower growth beginning in 1790 and continuing until 1840. Researchers stated, “This reduction in radial growth probably reflects a prolonged period of cooler summer temperatures that reduced the growing season at . . . high-altitude sites.”31 Further study of the Fremont Glacier concluded that in a colder climatic regime, as evident in the Little Ice Age, the climate might have been windier and drier.32 If the cooling period ended around 1840 –1850, as data seem to indicate, it would seem that

Environmental Change in the Red Desert

123

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 124

trappers arrived in the region when the climate still showed the effects of the Little Ice Age. Glacier studies in the Wind River Range place the end of the Little Ice Age at about 1845. The late stages of this period ushered in the present climatic regime and occurred at a time when Europeans first began to arrive in large numbers in the region.33 The ice cores from Upper Fremont Glacier indicate that in the mid-1800s a sudden climatic change took place.34 The glacial studies at the Upper Fremont Glacier and tree ring dates also show that the termination of the Little Ice Age was abrupt and occurred around ad 1845. This sharp change gave way to climatic patterns similar to today’s.35 Additionally, the archaeological excavations at Fort Bridger seem to support the evidence that the Little Ice Age ended in the early 1840s. In 1842, Fort Bridger sat on a hill not far from its present site. In 1843 the fort was moved to the valley floor. Before 1843 the valley would have been a wet meadow, but by 1843 the area had dried out to a point that the fort could be constructed in the valley. Jim Bridger’s fort was set directly on “gley soil,” a soil that indicates a wet meadow. It could not have been constructed there prior to the drying of the meadow. Interestingly, since 1843 there have been no over-bank episodes caused by flooding, so from that year to the present, the relocated Fort Bridger site has been relatively dry.36 The past environmental history shows that the weather in the Wyoming Basin has followed cycles, irregularly alternating between wet and dry and cold and not so cold.37 Soil erosion, often an indicator of shifts in the environment, seems to have accelerated in the last 200 years. The biggest change came not long after horses (ca. 1700 to 1750) and trappers (ca. 1820 to 1824) reached southwest Wyoming. It might be suggested that trapping-out beaver and a growing horse population that led to overgrazing played roles in causing soils to wash or blow away. Casual observers who travel through southwest Wyoming today and look at soils are struck by two prominent features: sand dunes, and arroyos or gullies. Dunes form under several conditions, but one simple way of 124

describing their formation is that they appear when there is little vegetation to hold sediments in place during windstorms. Sand particles collect anyplace there is a collecting feature like a hillside or rise in the land. The dunes that form move with the wind until enough moisture falls to allow plants to take root in the sand and stabilize the soils. Arroyos, on the other hand, are cuts into the land. They are formed by water action, generally caused by heavy runoff flowing over soils with sparse vegetation cover. This arroyo cutting is the direct result of increased seasonal flow. The water cuts progressively deeper into the soil until a more resistant layer of gravel or bedrock is reached. Once this occurs, the water begins to cut at the sides of the earthen canyon the runoff has created. This side cutting widens the arroyo, and over time the sediments in the valleys are scoured out. This downward and sideways cutting is natural in desert climates. Eventually a valley floor reaches a point of equilibrium, and the soils slowly build up until the next major downcutting cycle begins. There is much debate over how much Native Americans affected the environment of southwestern Wyoming. Native Americans set fires for several reasons: for opening grazing land for deer, as a weapon in warfare, or for camouflage. Accidents happened and hearth fires turned into range fires, but the effects of fires on the environment are questioned. Some believe that fires killed weak or infected plants. Others hold that fires removed ground cover and led to erosion. Another controversial question is whether or not Native Americans overhunted and killed off certain game species. Some scholars contend that Native Americans hunted certain Pleistocene mammals to extinction. Yet many species not only remained in the region after the end of the Ice Age, but thrived. Nature has provided one animal that slows arroyo cutting: beaver. Beaver did and can live in the narrow, deep streams that fed the Green River and Little Snake River. Carl P. Russell, in his classic work, Firearms, Traps, and Tools of the Mountain Men, remarks, “To the uninitiated, the large expanses of sagebrush coun-

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 125

try west of the Rockies may not appear to be wet enough to attract the water-loving beaver. In reality, however, the sage-filled basins are often laced by numerous small streams, and the small streams unite to form rivers. . . . In the earliest days of the mountain man some of these vast valleys held such concentrations of beaver as to encourage the hunter to kill his quarry with a club.”38 The beaver dams created ponds that trapped sediment and slowed the creation of arroyos. If Russell is right and beaver abounded in southwestern Wyoming, the landscape must have looked very different from today’s sage basin— creeks filled with beaver-created dams that held back water that, in turn, charged the water tables in the valleys. More trees and definitely more willows lined the creeks. More willows meant more beaver, and more beaver more ponds. Although the area was never a lush, treecovered wilderness, it was a desert with ribbon oases lining streams and creeks. What happened? Beaver dams and ponds still can be found in drainages like Currant Creek. One question is, how many beaver lived in the streams of the Wyoming Basin when Europeans arrived? What effect, if any, did beaver have on the environment? Robert Stuart was the first Euro-American to cross the northern end of the Wyoming Basin. On October 19, 1812, Stuart approached the west side of South Pass. As he headed east, he reached the Big Sandy. Not yet in the heart of the basin, he noted a considerable flat covered with willow.39 The next day, he crossed an alkali flat. He called the alkali “crystallized . . . salt.” The following date, he noted that his party did not find a drop of water until one of the men in his company “found a little bad water 11/2 miles from camp.” On the twenty-second he proceeded “east by south” over broken ground whose soil was principally sand and covered with little grass, some sage, and a good deal of salt wood.”40 On the twenty-third he described reaching a plain bound “only by the horizon” and had to “encamp without a single stick of wood.” Stuart did not take a direct route to the Sweetwater River. Instead of crossing through the heart of South Pass, he went in a di-

rection that took him across the northern edge of the Red Desert. One of his complaints was the lack of grass and water. South of Green Mountain he struck Lost Soldier Creek, which he called a “small branch with water.” Heading east Stuart reached Muddy Gap on the twenty-sixth of October. Turning north, he struck Muddy Creek, which he gladly noted had “muddy banks, and a great abundance of willow, running with a considerable body of water.” In this creek one of his party caught a beaver, and the next day they killed two buffalo bulls. Finally, on the twenty-eighth of October, he reached the Sweetwater River.41 Stuart had crossed not only South Pass but a stretch of desert that would become well known to tired travelers as a land of alkali, dust, and sage. What he did not note was arroyos. This omission is probably because of his route but may also be because arroyos were not there or not noteworthy. It would be nearly two decades before large numbers of trappers began to arrive in southwest Wyoming. James Clyman and company crossed South Pass in January or February of 1824. Clyman wrote, “We struck Sandy . . . there being a small growth of willows along the stream we had wood & water plenty.” Clyman reached the Green River on the twentieth of February. Here “Capt Smith with seven men” left the party to go farther south. Clyman wrote, “We left to trap on the branches of the stream as soon as the ice gave way . . . the ice giving way we found beaver plenty and we commenced trapping.”42 He provides an interesting discussion of the beaver: “We found a small family of diggers or Shoshone Indians on our trapping ground whom we feed with the overplus of Beaver the snow disappearing our ‘diggar friends’ moved off without our knowledge of when or where and when they had gone our horses running loose on[e] night they all disappeared and we were unable to find them or in what direction they had gone. We continued trapping on foot with fair success for about six weeks.”43 Clyman’s record also indicates the presence of willows along the Sandy. He did not specifically note the presence of beaver along this stream.

Environmental Change in the Red Desert

125

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 126

William Kittson traveled along the Bear River in 1824. While in the mountains that drain into Bear Lake, he reported on the exact number of beaver his company trapped. On May 18 he wrote that “trapps sett last night produced this morning 109 beaver.” On the nineteenth he and his men trapped sixty-eight more beaver, and the next day they moved camp. By the twenty-third, Kittson had reached the “American Territories.” Though not in the drainages of the Wyoming Basin, Kittson was on its fringes. His account provides one of the few records of the daily takes in beaver. There are not many firsthand perspectives of the Wyoming Basin from the 1820s, but William Ashley’s 1825 “Narrative” is one. From March 27 to April 1, 1825, William Henry Ashley trudged up and over the continental divide perhaps by way of the as yet unidentified Bridger Pass. On the twenty-eighth of March he made what may be one of the earliest references to arroyos, writing, “The country generally—is becoming cut to pieces by ravines or dry hollers some of the valleys are very fertile but their dryness renders them unfit for cultivation—snow is so common I have omitted to note its falling two days out of three.”44 The twenty-ninth of March took him closer and possibly over to the crest of the continental divide, west of present-day Rawlins.45 He goes on to describe the region he crossed heading west. They traveled “over ridges and valleys alternately, the latter the most extensive and generally covered with water produced by the melting of snow and which appeared to have no outlet.”46 His diary provides further information about these ponds: “This would be a difficult country to travel through at any other time than when the snow was melting as it would be entirely destitute of water entirely a bed of sand the only fuel small sage.”47 At this point he may have crossed into the Great Divide basin. His reference to the pond-dotted valleys and sage describes the general conditions in 1825. He added that “the dividing ridge is almost entirely destitute of vegetation except wild sage with which the earth is so bountifully spread that it proved a considerable impediment in our progress.” He may also have been the first Euro-American to 126

record crossing the Killpecker dune fields. On April 3 he “traveled about 12 miles west along the border of a ridge of sand that appears to continue for many miles b[ein]g e[ast] and w[est] 2 or 3 miles wide.” When he was camped near Steamboat Mountain, seventeen of his horses were stolen by “Indians.”48 While some of his party went off to look for the horses, Ashley headed down the flanks of the mountain into present Jack Morrow Creek and, angling west, eventually reached the Sandy River.49 Ashley’s description of the Sandy and a tributary that flowed into it parallels Clyman’s. Of the creek, Ashley wrote: “This stream is clothed with a growth of small willows and furnishes the only constant water we have met with since the 24th March and also the first wood we have seen in the same space.”50 He would follow “the meanders of the creek” to the Green River. He describes cottonwood, box elders, and willows growing along the banks of the Green and adds: “Judging from the quantity of the wood cut on its banks, and other appearances, it once must have contained a great number of beaver, the major part of which (as I have been informed) were trapped by men in service of the North West Company some four to five years ago.”51 If this account is correct, it would place trappers in the region as early as 1820 or 1821. Ashley described the soils from the Platte River to the Green River in general terms. The surface, he wrote, “generally either exhibits a bed of sand or a light brown barren earth, which is in many places wholly destitute of the least semblance of vegetation.” He felt that “in the summer and fall seasons of the year, the country will afford sufficient grass to subsist any number of horses in traversing it in either direction.” Ashley, unlike other earlier travelers, recognized the nutritious nature of the dry grasses in the area. He contended: “Such is the nutritious quality of the mountain grass that, when it can be had plentifully, although perfectly dry in appearance, horses (moderately used) that partake of it, will retain a great degree of flesh, strength and spirit.” Fifteen miles north of the confluence of the Sandy with the Green River, Ashley had a bull boat built, which he used to float his

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 127

goods down the Green River. In his diary he described Bitter Creek as “about 30 feet wide.” He then added that “this creek I called City creek from the appearance of the Country about its Junction.” He was referring to the palisade cliffs near present Green River, which gave the appearance of tall massed buildings.52 Possibly near the mouth of Currant Creek, he remarked on the considerable appearance of beaver and trapped four on the twentyseventh.53 Reaching the Henry’s Fork on April 30, 1825, he decided to make this the site of the first Rendezvous.54 On July 1, 1825, the Rendezvous would be held up the Henry’s Fork, according to Ashley’s reckoning, twenty miles from the Green River. Based on information Ashley gained at the Rendezvous, the Hudson Bay Company, over previous years, had trapped “one hundred and sixty thousand beaver” in the area claimed by the Shoshones. This would have included present southwest Wyoming. Ashley went on to say that “you can [gain from this] some idea of the quantity of beaver that country once possessed, when I tell you that some of our hunters had taken upwards to one hundred in the last spring hunt out of streams that had been trapped, as I am informed, every season for the last four years.”55 In 1830 Warren Ferris went west along much of Stuart’s route. At South Pass he descended into the Pacific Creek valley. On the twentieth of June he camped on the Sandy. This “small river,” according to Ferris, took its “name from the barren country through which it runs.” On the twenty-first, “after a weary march,” his party arrived at the Green River.56 Of the trip from South Pass to the Green River he wrote: “From the southern point of the Wind [River] Mountains, one or two snowy peaks rise, dimly visible to the southward—within the intervening space, a broken, sandy plain perfectly practicable for loaded wagons which may cross it without the least obstruction—separates the northern waters of the Platte, from those of the Colorado.” Ferris crossed the Green River on the twenty-sixth in a “bull boat” and “victimized” by mosquitoes.57 Ferris’s statement that the area could be crossed without obstruction bears comparison with the report of a much later trip.

In 1861 Mark Twain traveled by Overland Stage from the South Pass to the Green River along a similar route. At midnight the stagecoach ran into rain. The stage lost its way and wandered “about a plain with gaping gullies.” The “conductor” left the stage to walk ahead and chart a route back to the road. He immediately fell into “a chasm about fourteen feet deep,” and “his lantern followed like a meteor.”58 Twain’s colorful description of the arroyos indicates that erosion had occurred west of South Pass by 1861. Did Ferris overlook this erosion, or did it occur between 1830 and 1861? In 1834 John Kirk Townsend crossed South Pass and, like Ferris, followed the Sandy to the Green River. Townsend proved to be a keen observer. Crossing the pass and heading west, his party struck “the Sandy” on the fourteenth of June. He complained: “We found no water on the route, and not a single blade of grass for our horses.” Water and grass could be found, though not along the trail. The caravan on the fifteenth followed the course of the Sandy, where they found “a little poor pasture for the horses.” On the sixteenth, Townsend said he saw large herds of buffalo “on the plains of the Sandy River.” Possibly because he was following a trail taken by others who had grazed stock along the banks, he again remarked on finding no grass for his horses. When he arrived on the Green River, complaints about lack of grass for the horses stopped.59 Trappers and later travelers noted the general environment of the region in the 1830s. Sidney Smith, who traveled from Brown’s Hole to Ham’s Fork in June of 1839, described the area along the Green River north of the Uintas as a vast desert with “nothing to cheer up the path of the lonely traveler except the . . . croaking of the Ravens.” He probably did travel through the desert north of the Uintas, as he notes: “To the appearance there is no soil there is not a Spear of grass.”60 Smith in his travels in 1839 would have traveled over country close to the 1825 Rendezvous. The lack of grass could have been a factor of their 1839 route or could have reflected obvious changes to the region as early as the late 1830s. Beaver by the late 1830s became harder to find. Osborne Rus-

Environmental Change in the Red Desert

127

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 128

sell noted that along the Bear River he could find only beaver and antelope, and “Beaver were . . . getting very scarce.” Russell would later write: “The trappers often remarked . . . that it was time for the White man to leave the mountains as beaver and game had nearly disappeared.”61 Kit Carson, who spent the winter of 1841 in Brown’s Hole, put it this way: “Beaver was getting very scarce, and . . . it was necessary to try our hand at something else.”62 Traffic west along the California-Oregon-Mormon Trail and later the Overland Trail increased in the 1840s. In 1841 John Bidwell descended the west side of South Pass. In July, heading to the Green River, he noted the country along the Big Sandy to be “extremely dry and dusty” but made no mention of ravines, gullies, or arroyos.63 James John’s diary of 1841 shows how different opinions of the route west could be. For the same period, John notes in mid-July that he “camped on a creek of the Big Sandy. Here is good grass and the wild onions grow about 3 feet high.”64 In 1843 Kit Carson guided John Charles Fremont into the Wyoming Basin. On August 8 they crossed South Pass and began their trip down the Little Sandy the following day. Fremont noted that the creek flowed over a sandy bottom and had willows along its banks. On August 9 he reached the “Big Sandy.” He wrote that “the face of the country traversed was of brown sand of granite materials.” On the Big Sandy they turned north toward the west flanks of the Wind Rivers.65 There is little indication that there were deep arroyos along the route. Given Mark Twain’s account from the 1860s, when did the arroyos form along the Sandy?66 It is almost as if different travelers viewed a different landscape every year. Possibly they did. The fact is that in southwest Wyoming the cycles of moist and dry can come as yearly or seasonal events. These cycles of dry and wet can contribute to arroyo formation. If, in a dry year, the range is overgrazed, there is little vegetation to hold the heavy runoff from downpours the following season. Heavy wagon traffic along the Sandy may have led to overgrazing. The lack of food, and water that sometimes contained too much alkali, killed livestock. On July 20, 1849, Bernard Reed wrote that Little Sandy’s bottom reeked of the smell of dead 128

oxen.67 Four years later it was almost as if some of the travelers had traveled along another road and described a different landscape. Lorena Hays wrote in July of 1853, a few miles after the party left South Pass and came to Pacific Springs, “I have never seen a more beautiful road . . . we have encamped upon a large plain, surrounded mostly by what appears to be elevated, detached table lands, which in the smokey distance toward the setting sun, give an air of obscure and romantic caste to the scenery.” Hays noted that upon reaching the Big Sandy there was a “little green grassy bottom, the willow and a few other bushes giving quite a welcome.” Here she picked “currants.” Her tune changed as they traveled downstream and, she complained, “passed over the same parched sandy, grassless prairie” until they reached the Green River.68 Hays’s party might have journeyed over South Pass either before large numbers of travelers’ horses and oxen had eaten the grass, or come through in a year when more moisture reached the subsoil. The official journal of an 1856 handcart expedition noted that, along the Big Sandy, water and wood could be found, but not good feed.69 Horace Greeley wrote in 1859 that west of South Pass “no hills are visible — nothing but a sandy, barren plain, mainly covered with the miserable sage-brush.” Traveling in early July, he describes the Big Sandy as having a “bottom or interval of perhaps four rods in average width, in which a little grass is found. . . . Drouth and sterility reign here without rival.”70 To what degree beaver prevented or slowed arroyo formation in the Wyoming Basin before the arrival of Europeans has never been fully addressed.71 One reason is that the amount of moisture varies both from place to place and over time, because of differential rain and snowfall and the varied topography. The rapid removal of beaver via trapping also denies us a clear picture of what the region looked like prior to trapping. The impact humans had on the area is fraught with complexity. A single illustration explains how humans can affect the historical environment. At Rock Springs, water ran out of a spring under the rocks, the source of the town’s name, at a place on the northeast side of today’s College Hill. When the coal mines opened in the 1860s,

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 129

they cut deep into the strata under the “rock springs” and gravity began to draw the water away from the original spring vent and into the mine. The spring disappeared. Even if beaver had remained in the area, the loss of water would have prevented beaverdam building downstream from the spring. While Killpecker Creek, the stream that the spring once flowed into, may have had enough water to support beaver, it too lost fluid into the mine, reducing the creek’s volume at its confluence with Bitter Creek. All of this meant less water heading downstream through Killpecker Creek. Moisture falls in uneven amounts over the area. Today (2005) the annual precipitation in Green River is about eight inches per year,72 although some of the northern slopes of the higher ranges surrounding the basin still may get fourteen to twenty inches of effective moisture per year. Still, the question remains, when did arroyos appear? The process of erosion’s scouring valley floors and building up new soil deposits is as old as the desert west.73 The process continues to the present, but it is likely that in some places it was spurred on by specific factors that marked the end of the Little Ice Age. Along Bitter Creek, whose headwaters lie in an area that is drier than the Big Sandy’s, Stansbury noted arroyo erosion in 1850. He wrote, in September of that year, that from the confluence of Bitter Creek with the Green River the route east along the Bitter Creek valley was “intersected with gullies.” Not yet called arroyos, these “gullies” made the six miles up the valley “quite rough.”74 Captain Marcy, who traveled the valley in 1857, reported: “Bitter creek is a narrow and deep stream with high abrupt clay banks, making it very difficult to cross or approach with animals.”75 In 1863 Howard Cutting, on his trip west to the gold fields, found that from Rock Springs to Green River the road had to wind around endlessly to avoid the arroyo cuts. Cutting complained they had to “travel near three miles to make one.” 76 Union Pacific Railroad surveyors’ maps in 1867 and 1868 showed the arroyos, and A. J. Russell’s photographs illustrate the extent of the erosion in 1868. By the 1860s, erosion had cut the arroyos in lower Bitter Creek. When the downcutting took place is not yet

known. Probably multiple factors accelerated the erosion: the removal of beaver, the end of the Little Ice Age, increasing wagon traffic, and even overgrazing by horse herds likely accelerated the natural erosional process. Land can develop gullies rather quickly, especially in arid climates with low plant cover. Fires or overgrazing can remove the plant cover and start a swift process of erosion. Gullies were a maddening hindrance to travelers. Some, only a few yards wide, could force wagons to go miles out of their way. And deep, branching gullies could present the same difficulties as heavily crevassed ice in Arctic and Antarctic regions. The emigrant traffic, with attendant oxen, cows, and horses estimated in the thousands, ate what plant life they could find near the trail. The overgrazing reduced the plant cover, and when thunderstorms or floods struck, the water quickly cut into the unprotected soils, deepening and widening arroyos. By 1868 the arroyo near the mouth of Bitter Creek had been captured on film. A. J. Russell’s photograph of the Union Pacific grading camp at Carmichael’s cut, roughly three miles east of the Green River, captured the downcutting of Bitter Creek. Other creeks and streams had likewise begun to erode away the loose soils in the valleys. There is a saying that “only fools and newcomers try to predict Wyoming’s weather.” In many ways it is even harder to describe past climates. Geologists, archaeologists, and climatologists all have different ways of looking at the past. Most scientists would agree that the region is arid because of the rain-shadow effect caused by the Wasatch, western Wyoming, and Uinta mountains. But they do not agree on many other things except that the region is becoming drier and warmer. Even that is debated. What we do know is that humans have had an effect on the environment for a long time. Dams, ditches, roads, trails, and structures all altered the landscape. The built landscape was created at the cost of loosening soils that blew away and settled in dunes. When the beaver were removed in the nineteenth century, few could anticipate what would happen to the land by not having dams in small drainages. When roads were bladed and thin soils stripped to

Environmental Change in the Red Desert

129

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 130

bedrock, few considered the effect on future generations’ quality of life in the region. We are still not sure what the climate was like in the past—but we can guess that, just like the removal of the beaver in the nineteenth century, the removal of topsoils by development in the twenty-first century will accelerate erosion. And perhaps that is humankind’s major impact on the land; we vastly speed up the process of erosion, derailing the environmental processes that return the land to balance. When the next wet cycle occurs, the loosened soils will wash downstream because many of the hardy desert plants that hold back soil erosion will have been bladed away. In the future, scientists will debate the wet and dry cycles, but their view will be of a landscape with fewer native plants and more barren expanses.

130

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 131

Notes 1. A section of this tree is currently housed at Fort Bridger, Wyoming. When the military abandoned Fort Bridger in 1890, it appears that the tree was no longer irrigated. It is important to note that the nature of the environment since the end of the last ice age is still not well understood. As more archaeological work is done in the region, we will gain a better understanding of past climates. 2. The date humans arrived in southwestern Wyoming is constantly being pushed back in time as the result of new findings and excavations. The latest statewide radiocarbon assay files are located at “Wyoming Radiocarbon Dates,” Wyoming State Historic Preservation Office, http://wyoshpo.state.wy.us/shpoweb2002/2002webpages/c14dates.htm. The Blue Point site produced a bone collagen date from a horse (Equus) of 12,020 ± 40 years bp. This is a noncultural date (i.e., not related to human occupation). The Krmpotich site is a Folsom site and could push the occupation of the region back to the Folsom period (ca. 10,000 to 10,900 bp). See Michael R. Peterson, “Folsom Occupation and Technological Organization at the Krmpotich Site: An Analysis of the Lithic Assemblage” (master’s thesis, University of Wyoming, 2001), for a discussion of materials recovered at this site. George Frison, in Prehistoric Hunters of the High Plains (Academic Press, New York, 1978), 50, 26, dates the Folsom to 10,200 to 10,900 bp and notes that the Finley site in southwest Wyoming was occupied from about 8950 to 9100 bp. It is interesting to note that Pleistocene soil and sediment can still be found in the area’s ephemeral valleys. 3. William Eckerle, Rebecca Hanna, and Sasha Taddie, “Geoarchaeological Assessment of the Blue Point Site 48SW5734, Sweetwater County, Wyoming: Albert Cody Complex Adaptation to the Green River Basin Uplands,” in David E. Johnson and Jana Pastor, eds., The Blue Point Site: Paleoindian/Archaic Transition in Southwestern Wyoming (Western Archaeological Services, Rock Springs, Wyo., 2003), 1– 84. 4. Johnson and Pastor, Blue Point Site, Appendix A, 7. Eckerle notes that winters were no harsher than at present but “seasonality was much reduced, resulting in colder summers. Annual temperature in the study region was 12°C colder than present.” See also Jana Pastor, Kevin W. Thompson, Richard K. Talbot, William P. Eckerle, and Erik E. Ingbar, “Seeds-Kee-Dee”: Riverine Adaptation in Southwest Wyoming, Museum of Peoples and Cultures Technical Series, no. 99-3 (Brigham Young University, Provo, Utah, 2000), 417. Cf. Robert L. Kelly, Early Holocene Prehistory in the Intermontane West: Reinvestigation of the Pine Springs Site, Southwestern Wyoming, NSF 99-73272, Final Report (Department of An-

thropology, University of Wyoming, Laramie, 2001). 5. See Mark Miller, Paul Sanders, and Julie Francis, The Trappers Point Site (48SU1006): Early Archaic Adaptations in the Upper Green River Basin, Wyoming, vols. 1 and 2 (Office of the Wyoming State Archaeologist, University of Wyoming, Laramie, 1999), for discussion of the size of antelope during the Archaic period. Equus was found at the Blue Point site dating to 12,100 bp (Johnson and Pastor, Blue Point Site, 51). 6. See Kevin W. Thompson and Jana Pastor, People of the Sage: 10,000 Years of Occupation in Southwestern Wyoming, Cultural Resource Management Report 67 (Archaeological Services of Western Wyoming College, Rock Springs, 1995). For a discussion of environmental change in the region, see Johnson and Pastor, Blue Point Site, Appendix A, 7–18. In discussing weather changes of the past, sometimes scientists are looking at relative differences of a degree or two in temperature or an inch or two more of rain. Small differences can have profound influence on which plants grow in an area. 7. Tim Flannery, The Weather Makers (Text Publishing Co., Melbourne, 2005). Milutin Milankovich’s 1941 Canon of Insolation of the Ice Age Problem is today considered a breakthrough masterwork in climatology. 8. Pastor et al., “Seeds-Kee-Dee,” 417. 9. Johnson and Pastor, Blue Point Site, Appendix A, 7–18; Thompson and Pastor, People of the Sage. 10. William Eckerle, “Geoarchaeology of Altithermal Sand Dunes: Adaptation to Eolian Environments during the Early Plains Archaic” (master’s thesis, University of Wyoming, 1989), 28 – 30; Johnson and Pastor, Blue Point Site, Appendix A, 7–18. Eckerle notes at the Blue Point site the dunes began to form earlier than 12,100 bp. This comes from a collagen date from a horse bone recovered in Eolian deposits a meter above bedrock at the Blue Point site (Johnson and Pastor, Blue Point Site, 51, and see Appendix A, 7–18). 11. James Hendrick Mayer, “The Krmpotich Folsom Site (48SW9826): Geoarchaeology of a PaleoIndian Locality in the Killpecker Dunes of Southwestern Wyoming” (master’s thesis, University of Wisconsin, Madison, 2001), 64. 12. See Alan E. Wimer, “Folsom Sites in Southwestern Wyoming” (master’s thesis, University of Wyoming, 2001), 16 –17. Wimer notes that Jim Miller considers this increase in moisture to have lasted until 9500 bp. 13. Pastor et al., “Seeds-Kee-Dee,” 417. 14. Ibid.; Jane M. Beiswenger, “Late Quaternary Vegetational History of Grays Lake, Idaho,” Ecological Monographs 61, no. 2 (1991): 150 –162.

Environmental Change in the Red Desert

131

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 132

15. Mayer, “Dunes of Southwestern Wyoming,” 66. 16. Pastor et al., “Seeds-Kee-Dee,” 417, 419. Cf. David B. Madsen, “Environmental Change during the Pleistocene-Holocene Transition and Its Possible Impact on Human Populations,” in Charlotte Buck, ed., Models for the Millennium: Great Basin Anthropology Today (University of Utah Press, Salt Lake City, 1999), 75– 82. 17. Mayer, “Dunes of Southwestern Wyoming,” 69. In the Casper dune field, three to four meters of sand accumulated along the eastern margins at this time. See also Eckerle, “Geoarchaeology of Altithermal Sand Dunes.” 18. Mayer, “Dunes of Southwestern Wyoming,” 69. 19. Alan D. Reed and Michael D. Metcalf, Colorado Prehistory: A Context for the Northern Colorado River Basin (Colorado Council of Professional Archaeologists, Denver, 1999), 25. For the complexities associated with environmental change during this period, see Antony Thomas Plastino, “A Proposed Holocene Depositional Chronology, Sand Draw Basin, Jonah Gas Field, Sublette County, Wyoming, USA” (paper presented at the Seventh Biennial Rocky Mountain Anthropological Conference, September 16 –18, 2005, Park City, Utah). 20. Pastor et al., “Seeds-Kee-Dee,” 419. 21. Ibid. 22. Kenneth Lee Peterson, Climate and the Dolores River Anasazi: A Paleoenvironmental Reconstruction from a 10,000-Year Pollen Record, La Plata Mountains, Southwestern Colorado (University of Utah Press, Salt Lake City, 1988), 115. Here we use Peterson’s dates. 23. Ibid. 24. A. Dudley Gardner, Gabrielle Elliott, and Melissa Pola, “Granaries in the Douglas Creek Drainage Basin” (paper presented at the Colorado Council of Professional Archaeologists Conference, Grand Junction, Colorado, March 5, 2005), 1–10. 25. These data are just now being synthesized and much work needs to be done before drawing any conclusions, but it appears that the corncobs dating to 1200 bp (ad 750) to 1050 bp (ad 900) may dominate the assemblage. 26. Peterson, Paleoenvironmental Reconstruction, 115. From this point forward, unless otherwise noted. dates are expressed as ad, or as they appear in the cited source. 27. Pastor et al., “Seeds-Kee-Dee,” 420. 28. See Eric C. Carson, “Fluvial Response to Holocene Environmental Change, Uinta Mountains, Northeastern Utah” (Ph.D. diss., University of Wisconsin, Madison, 2003), for a discussion of glacial activity in the Uinta Mountains. Carson and Munroe note that, during the Little 132

Ice Age, glaciers did not grow. They write that “the data [suggest] that the Uinta Mountains experienced climate conditions different from those of surrounding mountains during the ‘Little Ice Age.’” They go on to say that “climatic conditions in the Uinta Mountains were sufficiently warm and/or dry to prevent glacier growth.” Eric C. Carson and Jeffrey S. Munroe, “Tree-Ring Based Streamflow Reconstruction for Ashley Creek, Northeastern Utah: Implications for Palaeohydrology of the Southern Uinta Mountains,” Holocene 15, no. 4 (2005): 609. 29. Mitchell A. Plummer, “Little Ice Age Climate in the Wind River Range, Wyoming, Estimated from Ice-Core Data,” Geological Society of America Abstracts with Programs 35, no. 6 (September 2003): 542. 30. See David L. Naftz, R. W. Kulsman, Robert L. Michel, Paul F. Schuster, M. M. Reddy, H. E. Taylor, T. M. Yanosky, and E. A. McConnaughey, “Little Ice Age Evidence from a South-Central North American Ice Core, USA,” Arctic and Alpine Research 28, no. 1 (1996): 35– 41; Paul F. Schuster, D. P. Krabbenhoft, David L. Naftz, L. Dewayne Cecil, Michael L. Olson, J. F. Dewild, David D. Susong, J. R. Green, and M. L. Abbott, “Atmospheric Mercury Deposition during the Last 270 Years: A Glacial Ice Core Record of Natural and Anthropogenic Sources,” Environmental Science and Technology 36, no. 11 (2002): 2303–2310; Paul F. Schuster, David E. White, David L. Naftz, and L. Dewayne Cecil, “Chronological Refinement of an Ice Core Record at Upper Fremont Glacier in South Central North America,” Journal of Geophysical Research 105, no. D4 (February 2000): 4657– 4666; David L. Naftz, D. D. Susong, P. F. Schuster, L. D. Cecil, M. D. Dettinger, R. L. Michel, and C. Kendall, “Ice Core Evidence of Rapid Air Temperature Increases since 1960 in Alpine Areas of the Wind River Range, Wyoming, United States,” Journal of Geophysical Research 107, no. D13 (2002): acl 3-1 to acl 3-16. 31. Naftz et al., “Little Ice Age,” 39. Naftz et al. do mention one warming period during the Little Ice Age, dating to the mid-1700s. They note that tree ring studies for the west indicate a warming trend from ad 1650 to 1740. 32. Schuster et al., “Chronological Refinement,” 4664. Schuster et al. offer the following: “A colder climatic period, such as the Little Ice Age (LIA), will typically be windier and dryer. . . . Warmer climatic periods are typically wetter and less windy“ (4664). 33. Ibid., 4657– 4666. This article specifies that most of the “ice deposited at the bottom of the glacier formed from snows which fell between A.D. 1716 and 1820” (ibid., 4657). 34. Naftz et al., “Little Ice Age,” 35. Naftz used grasshopper legs to obtain a radiocarbon assay that allowed him to date the Little Ice Age snow deposition to 1716 –1820. See ibid., 37, for the calibration of the radio-

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 133

carbon dates used in his study. Schuster et al., “Chronological Refinement,” contend that there might have been postdepositional movement of the grasshopper legs, so that there was a possible dating error of 95 years (4659). Schuster et al. used volcanic fallout as stratigraphic indicators. In the nineteenth century, two explosions—Tambora, March 7– 12, 1815, and Krakatau, August 26 –27, 1883—proved to have left usable chemical signatures in the Fremont Glacier (4462). Table 1 in their work notes thirty-four major volcanic eruptions in the last 250 years that could have potentially left fallout on the Upper Fremont Glacier surface (4663). They did find thirty suspected volcanic fallout layers, indicating that “global volcanic activity was . . . high . . . during this period” (4660). They wrote: “Using the known age dates (3H, 36Cl, 14C, Krakatau and Tambora volcanic events), additional historical volcanic events were assigned to ECM [electrical conductivity measurement] peaks in the ice core and used as additional time-line events to calculate a refined agedepth profile. . . . At a depth of 152m the refined profile also shows good agreement (1736 A.D.) with the 14C age date (1729 A.D. ±95 years)” (4665). Schuster et al. put the end of the Little Ice Age at around 1845 (4657– 4666). 35. Schuster et al., “Chronological Refinement” (4664– 4665). 36. William Eckerle, “Geological, Pedological, and Paleoenvironmental Assessment of the Bridger Trading Post Locality, Fort Bridger (48UT29), Wyoming,” in A. Dudley Gardner, David E. Johnson, and Debbie Allen, eds., Data Sets (1992) and Paleoenvironmental Assessment (1993) for Fort Bridger (Western Wyoming Community College and Archaeological Services, Rock Springs, 1994). The first section of this report provides the historical context. 37. It should be noted that not all authorities agree with this cycle of climatic change. Cf. Jim Miller, “Geology in Archaeology: Geology, Paleoclimates, and Archaeology in the Western Wyoming Basin” (master’s thesis, University of Wyoming, 1992). Miller does agree, however, with many of the broader climatic themes presented in Table 1 of Schuster et al., “Chronological Refinement.” Miller clearly notes that arroyo cutting is due to increased seasonal flow. 38. Carl P. Russell, Firearms, Traps, and Tools of the Mountain Men (University of New Mexico Press, Albuquerque, 1967), 146. For a discussion of what happened to Pleistocene mammals, see Robert L. Kelly and M. Prasciunas, “Did the Ancestors of Native Americans Cause Animal Extinctions in Late Pleistocene North America?” in M. E. Harkin and D. R. Lewis, eds., Reconsidering the Ecological Indian (University of Nebraska Press, Lincoln, in press). 39. Phillip Ashton Rollins, ed., The Discovery of the Oregon Trail: Robert

Stuart’s Narratives of His Overland Trip Eastward from Astoria in 1812 –13 (University of Nebraska Press, Lincoln, 1995), 162. 40. Ibid., 162 –164. 41. Ibid., 165–166. 42. James Clyman, Journal of a Mountain Man (Mountain Press, Missoula, 1984), 29 –30. The editor of this volume, Linda M. Hasseltrom, puts the year at 1824 (29). 43. Ibid., 30. Cf. George C. Frison, “Shoshonean Antelope Procurement in the Upper Green River Basin, Wyoming, Plains Anthropologist 16, no. 54 (1971): 258 –284. 44. Dale L. Morgan, ed., The West of William H. Ashley: The International Struggle for the Fur Trade of the Missouri, the Rocky Mountains, and the Columbia, with Explorations beyond the Continental Divide, Recorded in the Diaries and Letters of William H. Ashley and His Contemporaries, 1822 – 1838 (Old West Publishing Co., Denver, 1964), 104. Some of the original spellings have been changed in this publication— e.g., “vallies” to “valleys.” 45. Ibid., 104, 265. 46. Harrison Clifford Dale, ed., The Ashley-Smith Explorations and the Discovery of a Central Route to the Pacific, 1822 –1829 (Arthur H. Clark, Cleveland, 1918), 132 –133; hereafter cited as “Ashley-Smith Explorations.” 47. Morgan, William H. Ashley, 104. 48. Ibid., 104, 265. Initially Ashley blamed the “Snake Indians,” but it appears the horses were stolen by Crows. See ibid., 265, n87, for Morgan’s discussion of Jack Morrow Creek. 49. Ashley-Smith Explorations, 132 –133. The footnotes of this journal provide some confusion as to where precisely Ashley crossed the continental divide in 1824. It is assumed he crossed between Rawlins Spring and Separation Creek. If he went over Bridger’s Pass, to the south, he would have descended Muddy Creek. At the point where Muddy Creek bends south toward the Little Snake River, there is a broad valley where water stands in the spring. Ashley’s description, however, leads one to believe water stood in the numerous valleys through which he traveled. Morgan puts forward the idea that Ashley went along the future route of the Union Pacific railroad. Morgan, William H. Ashley, 104, 265. 50. Ashley-Smith Explorations, 135. Ashley’s “Narrative” is vague enough to create some doubt as to his precise location. He did, however, on April 12, head southwest to the Green River. On April 24, Ashley was on a tributary of the North Platte and may have reached the river that day. 51. Ibid., 136. 52. Morgan, William H. Ashley, 107. 53. Ibid.

Environmental Change in the Red Desert

133

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 134

54. Ashley-Smith Explorations, 139 –140. The site of the Rendezvous had to be moved up Henry’s Fork because of mosquitoes, though in April, when Ashley had selected the first Rendezvous site, the insects had not yet hatched. 55. Ibid., 157–158. The citations regarding Ashley are from his “Narrative.” 56. Warren Angus Ferris, Life in the Rocky Mountains: A Diary of Wandering on the Rivers Missouri, Columbia, and Colorado, 1830 –1835 (Old West Publishing Co., Denver, 1983), 111. 57. Ibid., 112. 58. Mark Twain, Roughing It (New York: Signet Classic, 1962), 88 – 89. 59. John Kirk Townsend, Across the Rocky Mountains to the Columbia (University of Nebraska Press, Lincoln, 1978), 77– 84. 60. Leroy and Ann Hafen, eds., To the Rockies and Oregon 1839 –1842: With Diaries and Accounts by Sidney Smith, Amos Cook, Joseph Holman, E. Willard Smith, Francis Fletcher, Joseph Williams, Obadiah Oakley, Robert Shortess, and T. J. Farnham (Arthur H. Clark, Glendale, Calif., 1955), 78 – 79. 61. Osborne Russell, Journal of a Trapper: 1834 –1843 (University of Nebraska Press, Lincoln, 1965), 112, 123. 62. Milo Milton Quaife, ed., Kit Carson’s Autobiography (University of Nebraska Press, Lincoln, n.d.), 63. The Bison Book edition of this book by the University of Nebraska Press notes the book first appeared as “No. 33 in the Lakeside Classics published in 1935, and is reprinted by arrangement with the Lakeside Press, R. R. Donnelley & Sons Co., Chicago.” 63. Dr. Doyce B. Nunis Jr., ed., The Biddle-Bartleson Party: 1841 California Emigrant Adventure (Western Tanager Press, Santa Cruz, Calif., 1991), 36 –37. 64. Ibid. 65. Donald Jackson and Mary Lee Spence, The Expeditions of John Charles Fremont, vol. 1, Travels from 1838 to 1844 (University of Illinois Press, Urbana, 1970), 254. 66. In 1861, when his brother Orin was appointed secretary to the Nevada Territory, the two brothers took the Overland Stage to Nevada. Twain would have taken the Overland Stage route over South Pass and down the Sandy toward the Green River and ultimately Fort Bridger. When Ben Holladay purchased the Central Overland California and Pike’s Peak Express Company in 1862, he quickly decided to move the mail and stage services south to the route Stansbury had explored. Mark Twain would not have traveled the Bitter Creek route on his return to the east in the winter of 1866 and 1867, as he traveled via steamships and a passage 134

across Nicaragua. These travel sketches were published in the San Francisco Alta California, December 14, 15, 1866; January 18, 1867; February 27, 1867; March 16, 1867. In terms of arroyos, Howard Stansbury provided a detailed description of the arroyos in the lower Bitter Creek valley in 1850. Stansbury noted on September 14, 1850, that the lower reaches of Bitter Creek were cut by “gullies” and that “the ground is sandy and is cut down perpendicular on each bank by the stream very deep (say 20 ft).” Brigham D. Madsen, Exploring the Great Salt Lake: The Stansbury Expedition of 1849 –50 (University of Utah Press, Salt Lake City, 1989), 631. 67. Mary McDougall Gordon, ed., Overland to California with the Pioneer Line: The Gold Rush Diary of Bernard J. Reid (Stanford University Press, Stanford, Calif., 1983), 88 – 89. 68. Jeanne Hamilton Watson, ed., To the Land of Gold and Wickedness: The 1848 –59 Diary of Lorena L. Hays (Patrice Press, Saint Louis, 1990), 177–179. 69. Leroy R. Hafen and Ann W. Hafen, Handcarts to Zion: 1856 –1860 (Arthur H. Clark, Glendale, Calif., 1981), 212. 70. Horace Greeley, An Overland Journey from New York to San Francisco in the Summer of 1859 (University of Nebraska Press, Lincoln, 1999), 193–194. 71. Rich Olsen and Wayne Hubert, Beaver: Water Resources and Riparian Habitat Manager (University of Wyoming, Laramie, 1994), 1–3, note that beaver dams elevate water tables, enhance riparian vegetation, and “trap eroded silt from adjacent land.” It is believed that before the arrival of Europeans beaver ranged from the arctic tundra to the deserts of northern Mexico. Stephen H. Jenkins and Peter E. Busher, “Castor canadensis,” Mammalian Species 120 (June 1979): 1– 8. The beaver population at the arrival of Europeans, according to Olsen and Hubert, stood between 60 to 400 million animals in North America. This would be the equivalent of roughly 10 to 60 animals per mile of stream. Robert J. Naiman, Jerry M. Melillo, and John E. Hobbie, “Ecosystem Alteration of Boreal Forest Streams by Beaver (Castor canadensis),” Ecology 67, no. 5 (1986): 1254–1269. Currently in Wyoming, beaver live in only about one-third of their original range. The dramatic decrease in the beaver population was the direct result of overtrapping. This decrease led to the drop in water tables, the loss of dams as sediment traps, and the loss of riparian vegetation. 72. Green River Resource Area Resource Management Plan (RMP) and Final Environmental Impact Statement, vol. 1 (U.S. Department of the Interior, Bureau of Land Management, Green River Resource Area, Rock Springs District, 1996), 313. The RMP notes, “Precipitation and tem-

Dudley Gardner

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 135

perature are the two most important climatic parameters relating to vegetation growth” (313). The plan also notes the significance of evapotranspiration on the environment, i.e., moisture loss due to wind (pan evaporation in the report is one measure used to determine evapotranspiration). The RMP states: “The estimated pan evaporation for the area is 65 to 70 inches of water. Pan evaporation indicates the amount of moisture that can be lost by water bodies, soils, and vegetation due to the atmospheric conditions” (313). 73. See Charles Ferguson, “Geology of the Red Desert,” Chapter 1 of this volume. 74. Madsen, Exploring the Great Salt Lake, 631. C. F. James Miller contends that downcutting is caused by numerous small-scale events, not big thunder boomers. James Miller, “Late Pleistocene and Holocene Geology of the Central Green River Basin with Reference to Geoarchaeology and ‘Benchmark’ Site Selection,” manuscript on file at Kemmerer Resource Area of the Bureau of Land Management, May 17, 1996. 75. R. B. Marcy to Major F. J. Porter, in The Executive Documents, Printed by Order of the Senate of the United States, Second Session, ThirtyFifth Congress, 1858 –59, and Special Session of the Senate of 1859, vol. 2, no. 1, part 2, Report of Secretary of War, part 1 (William A. Harris, Printer, Washington, D.C., 1859), June 28, 1858, 110 –111. 76. A. Dudley Gardner and Val Brinkerhoff, Historical Images of Sweetwater County (Donning, Virginia Beach, 1993), 35.

Environmental Change in the Red Desert

135

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 136

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 137

4. Titanotheres, Time, and People A Snapshot of Red Desert Paleontology Tom Rea

ill Reed was a little guy with a big moustache and an expansive sense of his place in the world. In the years before 1900, he was the most productive fossil collector in Wyoming. He began collecting as early as 1877, when he and a fellow railroad worker found dinosaur bones strewn about like logs on the north flank of Como Bluff, near Medicine Bow, east of the Red Desert. He went on to work for several years for the great paleontologist O. C. Marsh of Yale, during the years when Marsh’s competition with E. D. Cope of Philadelphia was at its most intense. By the mid-nineties Reed was collecting for Professor Wilbur Knight at the University of Wyoming, and by 1898 they had amassed a collection of Jurassic vertebrate fossils that was second only to Marsh’s at Yale. Bones were crammed in closets and corners all through the university’s two buildings. In December 1898 the New York Journal and Advertiser ran a full-page story under a big headline: “Most Colossal Animal Ever on Earth Just Found Out West.” In one corner was a photo of Reed standing next to a dinosaur femur taller than he was. He was incorrectly identified as “Reeder,” a UW professor—he was not a professor, though he knew as much as anyone about Wyoming’s rocks and fossils—the man who’d found “Brontosaurus giganteus.” In New York the steel tycoon Andrew Carnegie read the newspaper story. Before long, Reed had gone to work with paleontolo-

B

gists from the Carnegie Museum in Pittsburgh, charged with bringing back the big Wyoming dinosaur. The so-called Brontosaurus giganteus turned out not to be at the quarry where Reed thought he had left it; Knight or his students appear to have gotten back there first and taken or smashed the bones. Still, under Reed’s guidance, the Pittsburghers found another big dinosaur near Sheep Creek, about twenty-five miles north of Medicine Bow. This dinosaur was named a new species, Diplodocus carnegii, by the Carnegie paleontologists. Reed spent the winter of 1899 –1900 in Pittsburgh, working at the museum on the newly quarried bones. Perhaps to impress his new employers, he told them that he’d found and cached the skulls of five mammals from the Tertiary—that is, post-dinosaur times—in the Red Desert way back in 1885. He let them believe that finding the skulls again would be simple. In April 1900, with a Carnegie assistant named Olaf Peterson, Reed headed out from Rawlins for Picket Lake. Picket Lake lies just barely on the Sweetwater side of the divide along the north rim of the Great Divide basin, a few miles east of Honeycomb Buttes. Intermittent Picket Creek flows nearby. Nothing flows out of the lake. It’s a treeless place, with sagebrush in very sandy soil stretching in all directions. The lake covers perhaps 350 acres when full, 137

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 138

but by early September much of it is so shallow that the bottom mud spackles the water surface. The two Oregon Buttes rise on the western horizon, the Wind River Mountains hulk up to the northwest, with Steamboat Mountain far to the south. At first it seems surprising that this remote place would be well known to a bone prospector in 1885. But the lake is only a few miles south of the Seminoe Cutoff, one of several CaliforniaOregon-Mormon Trail braids that followed the Sweetwater up to South Pass. As a water source, even stagnant, brackish water, the lake would have been known to white travelers since the 1840s. Organized paleontology first came to the Red Desert with the Union Pacific Railroad. Cope, Marsh, and Joseph Leidy all collected in Eocene badlands around Fort Bridger, near the southwestern corner of the Green River basin, in the early 1870s. Leidy found and named the genus Uintatherium there—a big, rhinolike animal with bony protuberances on the front of its skull. At the same time, Cope and Marsh were scrambling to outpublish each other, naming other big mammals from the Eocene. Cope explored farther afield than the other two—along Ham’s Fork to the north, into Eocene strata in the Washakie basin a hundred miles east of Fort Bridger, and south of the railroad.1 By 1900 the discipline was beginning to mature, from competition and secretiveness toward a cooler-headed curiosity about the history of life. But Reed had come into the fossil business when the trophy hunting was in full swing, and he still thought more like a prospector than a scientist. That his employers also were intrigued by his story of the five skulls shows that they, too, were still susceptible to the fever of the treasure hunt. Reed’s correspondence doesn’t say what kind of skulls they were—perhaps Eocene titanotheres, bigger even than the uintatheres, or perhaps big cats from the Miocene. He may not have known. In any case, separated from where they’d first been found, the skulls would tell little about the life and times of the creatures they once were part of. They’d be useful only for museum filling, for display. Since Reed’s time, the Red Desert has told a great many stories about the history of life. Fossils include everything from 138

spiral-shelled ammonites to huge mammals, and what Jay Lillegraven, professor emeritus of paleontology at the University of Wyoming, describes as “jillions of turtles.” Fossil plants are plentiful too, among them palm trees and broad-leaved deciduous plants. The oldest widespread fossil material in the Red Desert dates from the mid-Cretaceous, around 100 million years ago, when a shallow sea bifurcated North America. The sea moved in and out of Wyoming numerous times as it rose and fell. Plenty of fossil oyster beds were left behind, and a great diversity of bottomdwelling clams and snails. The best Cretaceous exposures are in the Rock Springs Uplift, an oval, dome-shaped feature with its western edge at Rock Springs and its eastern edge near the Bitter Creek exit on I-80, about thirty-five miles east. The uplifted layers are like a set of nested bowls, but upside down, with their bottoms mostly worn off, so that the oldest layer — the midCretaceous Baxter shale—is exposed in the middle and sequentially younger layers appear as one moves outward from the center. In the older-to-younger sequence of these Cretaceous marine rocks, the baculites—squid relatives—are the signature fossils paleontologists use to date the layers. Toward the end of the Cretaceous, the sea receded and fossils of land animals begin to appear. These include a spectacular diversity of dinosaurs, crocodiles, and even mammals. Some are shrew-size, some mouse-size, some rat-size. A few got as big as badgers, but none were bigger than that before the extinction of the dinosaurs marked the end of the Cretaceous, 65 million years ago. After the Cretaceous came the Tertiary, which is divided into the Paleocene epoch starting 65 million years ago, then the Eocene, Oligocene, Miocene, and Pliocene, which ended about 2 million years ago when the ice ages began. The Paleocene is a particularly fascinating time in the history of life; not only were the dinosaurs suddenly gone, but so, too, were many of the mammals. “There were a lot of empty niches at the old corral,” says Lillegraven. Recovery was slow at first. Then

Tom Rea

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 139

it took off. Animals began evolving rapidly. Diversity was higher at first among small forms, but as animals steadily got bigger, diversity increased among the larger forms. In the Eocene, animals again changed dramatically. Virtually the entire first half of the Eocene—the first 10 million of its 18 million years—is represented in the Red Desert, and the desert is one of the best places in the world to find fossils from that time. A journey leaving I-80 just east of Point of Rocks and heading south to Adobe Town in the Washakie basin crosses all 10 million of those Eocene years in order, from old to young. “An absolute paleontological wonderland,” Lillegraven calls it. It was a time of freshwater lakes that grew and shrank, grew and shrank, and of rivers that ran between and among them. The lakes must have been spectacular, full of fishes, aquatic birds, and those jillions of turtles. The layers preserve the eons of complicated intertonguing among the lake beds and the fluvial rocks laid down by the changing rivers. On dry land, too, there was amazing faunal diversity, from tiny mammals to titanotheres at Adobe Town that would dwarf a living rhino—taller than a man at the shoulder. Then comes a big gap. Everything deposited for the rest of the Eocene, and for all of the Oligocene and most of the Miocene, is gone—blown or washed away. By the late Miocene there was a lot of volcanic activity in what is now Nevada and southern Oregon. Ash blew into the air there and fell out of the sky here, where it was reworked by wind and streams into rock and further weathered. By the late Miocene, predecessors of all the fauna we know now in North America, including those of pronghorn antelopes and many kinds of elephants, were here. A few, like horses and camels, started here and later migrated to Asia. Some of the oldest known horses are from the Red Desert—Hyracotherium, the little dog-size creature formerly known as Eohippus.2 Miocene ash beds and their fossils cross into Colorado from the southern edge of the Washakie basin. There are Miocene rocks too on the northern edge of the Red Desert near, among other places, Picket Lake.

As for Reed and Peterson, they spent two weeks in 1900 looking for the five skulls—perhaps in the Eocene exposures in the Honeycomb Buttes a few miles west of the lake, or in similar rocks to the northeast. Then Reed recalled that when he was at Picket Lake the first time years before, he’d only been hunting stray horses. In fact, he said, it was in central Wyoming north of the Rattlesnake Hills that he’d had such great fossil-hunting luck. Prospecting was so good there, he told Peterson, that he “picked up half a bushel of small skulls in half an hour.” But when the two men crossed the Sweetwater and traveled on north of the Rattlesnakes, they couldn’t find much there either. Peterson, disgusted, headed by himself for Tertiary fossil beds on Henry’s Fork, far to the southwest. Reed, after a few weeks more, found himself out of a job.3 For a hundred years, Reed was remembered at the Carnegie Museum as a liar and a blowhard. In Wyoming his reputation was more respectable. When Wilbur Knight died unexpectedly in 1903, Reed got his job teaching geology at the University of Wyoming. His students remembered him fondly. After he died, they put up a plaque to “this faithful and useful man of science” in the university geology museum, where it may still be seen today. Lillegraven doubts Reed’s five skulls ever existed. At least one other paleontologist, Chris Beard of the Carnegie Museum, is willing to give Reed the benefit of the doubt, if only because it’s at least possible Reed did find and cache some skulls in 1885 and later simply forgot where he left them. It hardly matters. More interesting is the slight irony that for nearly fifteen years now the Carnegie Museum has again maintained a paleontological presence in the Red Desert and is making some significant contributions to our understanding of the history of life. Beard, a 2000 MacArthur fellow, specializes in the changes mammals were undergoing in the Paleocene, when many new species of animals, plants, and even insects emerged. This is the same period Lillegraven found so interesting—with all its evolu-

Titanotheres, Time, and People

139

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 140

tionary opportunities after the dinosaurs disappeared. Beard is especially absorbed by how populations changed as they moved across continents; his work in Wyoming is bound up with Asian connections. He has worked in China and Tibet, and evidence he and his colleagues have found relocates our earliest anthropoid ancestors—the common ancestors of monkeys, apes, and people—from Africa to China. In Wyoming, Beard has been digging in three sites in the Paleocene Fort Union Formation since the early 1990s. One dig lies south and one north of I-80 along the eastern edge of the Rock Springs Uplift. A third lies at Bison Basin, on the northern edge of the Great Divide basin about sixteen miles northeast of Picket Lake. The world warmed throughout the Paleocene, the speed of the warming increasing dramatically as the epoch neared its end. These changes are well documented in a variety of sites in the Red Desert and elsewhere. Beard is especially enthusiastic about his sites in the Fort Union Formation, however, because they preserve fossil vertebrates and fossil plants together. Paleontologists can study plant and animal evolution simultaneously. At the same time, the Bering land bridge became more passable. Plants, animals, and perhaps insects kept arriving from Asia, and North American plants and animals came to look more like Asian ones. The invaders generally did well, Beard says, while the natives suffered.4 And while scientists like Beard are pursuing big questions like global climate change and intercontinental migration in their work, more modest kinds of curiosity, not so different from Reed’s, survive as well in the Red Desert. J. P. Cavigelli, a skilled fossil collector and preparator, is manager of the fossil prep lab at the Tate Geological Museum at Casper College. He maintains a limestone dig on private land in the Wasatch Formation near Bitter Creek, on the northwest edge of the Washakie basin. The limestone once was soft mud at the bottom of a freshwater pond. At the moment he’s working on a pile of fossil crocodile bones, which are not articulated, that is, not laid out in their original connec140

tions, but are still associated, that is, together in one place. Also on hand are crocodile scutes—their bony plates—and snails, and big scales from garfish.5 How remarkable that such dry, sandy environs were once so lush and moist. Work like Cavigelli’s helps us remember how entirely different the planet can be from what it is now. And that’s an important thing.

Tom Rea

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 141

Notes 1. David Rains Wallace, The Bonehunters’ Revenge (Boston: Houghton Mifflin, 1999), 77–91, 112 –130. 2. Jason Lillegraven, pers. comm., July 25, and e-mail messages, August 2, 3, 8, and 14, 2006. 3. Tom Rea, Bone Wars: The Excavation and Celebrity of Andrew Carnegie’s Dinosaur (Pittsburgh: University of Pittsburgh Press, 2001), 137–143. 4. Christopher Beard, pers. comm., August 9, 2006. 5. J. P. Cavigelli, pers. comm., August 10, 2006.

Titanotheres, Time, and People

141

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 142

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 143

5. Vertebrate Wildlife of the Red Desert Gary P. Beauvais

ildlife is central to the concept and attraction of wild land. It is difficult to define wilderness, semiwilderness, or perhaps even “countryside” without referring to free-ranging, native animals pursuing their life activities with minimal interference or assistance from people. The Red Desert is no exception. Much of its significance as a place derives from its wildlife and from the fact that this wildlife assemblage has remained relatively undisturbed by virtue of the desert’s high, dry, and cold setting. But while these conditions preclude disturbances like farming and urbanization, they do not preclude the extraction of oil, natural gas, and other minerals. So now the stage is set in a familiar manner: wild land and wildlife, with their tangle of tangible and intangible values, are pitted against mineral substrates that are quickly translated into dollars in an energy-hungry global market. If this situation is settled solely by short-term economics, the fate of Red Desert wildlife is predictably dark. This essay is a small attempt to inform, and encourage, other value systems that might be brought to bear. The fauna of any area cannot be completely understood without understanding that area’s spatial and temporal contexts.1 As much as it pains some theoretical ecologists to admit, faunas are influenced more by geographic surroundings and history than by

W

local or recent phenomena. The continental setting of the Red Desert determines the area’s physical conditions, and those conditions in turn favor animals with particular adaptations. Similarly, past events (some natural, some human-caused) have affected certain animals more than others. Physical Setting and Conditions A study of Red Desert wildlife is largely a study of adaptations to physical extremes. This place demands of its wildlife the ability to tolerate extreme aridity and cold, often simultaneously. The Red Desert lies in a triple rain shadow that permits only about ten inches of annual precipitation.2 Cold fronts from the northern Pacific have lost most of their water to the Cascades, the Sierra Nevada, and the Central Rocky Mountains by the time they arrive at the Red Desert. Similarly, moisture from the Gulf of Mexico (an important source of summer precipitation on the Great Plains) is blocked by the front ranges of the Rocky Mountains immediately to the east. Finally, the Arizona monsoon from the southern Pacific rarely makes it to the Red Desert in late summer, being all but completely wrung out by the Wasatch and Uinta mountains of Utah and the Southern Rocky Mountains of Colorado. 143

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 144

Table 5.1. Examples of peripheral vertebrates occupying the Red Desert Class

Species common name

Species Latin name

Ecoregion at core of range

Fish

Cutthroat trout Flannelmouth sucker

Oncorhynchus clarkii Catostomus latipinnis

Rocky Mountains Colorado Plateau

Amphibian

Great Basin spadefoot

Spea intermontana

Great Basin

Reptile

Western hognose snake Ornate tree lizard

Heterodon nasicus Urosaurus ornatus

Great Plains Colorado Plateau

Mammal

Dusky shrew Swift fox Pygmy rabbit

Sorex monticolus Vulpes velox Brachylagus idahoensis

Rocky Mountains Great Plains Great Basin

Bird

McCown’s longspur Gray vireo

Calcarius mccownii Vireo vicinior

Great Plains Colorado Plateau

Note: These are species that are native to portions of the Red Desert but whose continental ranges are centered in adjacent ecoregions.

Coupled with this low precipitation are the short growing season (about 147 frost-free days per year) and cold temperatures (average minimum January temperature about 8°F) typical of elevations between 4,900 and 9,500 feet. No other region of at least this size in the contiguous United States has as extreme a combination of aridity and cold. But lack of precipitation is only part of the reason why the Red Desert is so dry, and low temperatures are only part of the reason why it is so cold. Strong and persistent winds amplify aridity by accelerating evaporation of liquid water, sublimation of frozen water, and physical movement of fallen snow out of the region. The same winds also lower apparent temperatures for endothermic wildlife (“wind chill factor”) and quickly sap any solar heat accumulated in soil and dark rock.3 As with precipitation, the wind in the Red Desert is largely a function of surrounding landscapes. If the Rocky Mountains are a check dam against airflow across the continent, the Red Desert is a major breach. Air masses rush east between the Central and Southern Rocky Mountains at 144

this gap with enough consistency and ferocity to have built the second-largest sand dune field in the world—Killpecker Dunes. wildlife influences of adjacent region The Red Desert lies at the junction of four major ecoregions: Great Basin, Colorado Plateau, Rocky Mountains, and Great Plains. It is therefore not surprising to find in the Red Desert several species more characteristic of these provinces (Table 5.1). The desert is somewhat of a barrier, in that it allows some species from adjacent regions to enter but not pass through. The Red Desert has prevented several forest-adapted mammals from the Central Rocky Mountains from colonizing the Southern Rocky Mountains, and has also promoted the evolution of unique mammalian subspecies in each mountain province.4 Even some open-country mammals find the Red Desert a barrier. Swift foxes from the Great Plains occasionally range into the eastern Red Desert, and kit foxes from the Colorado Plateau ap-

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 145

proach the southern boundary (for the Latin names of all wildlife species mentioned in the text, see Table 5.2).5 Each species tolerates rather cold and dry conditions within its respective range— the swift fox in the frigid plains of Saskatchewan, and the kit fox in the Sonoran Desert, with less than four inches of annual precipitation. Yet neither has successfully traversed the Red Desert.

Just as it is not surprising to find peripheral species edging into the Red Desert, it is also not surprising to find some widespread species in the area (Table 5.3). Some display ecoregional affinities at the subspecies level. Although gopher snakes occur across the western United States and much of the Red Desert, the Great Plains subspecies occupies the eastern side of the

Table 5.2. Wildlife species mentioned in this chapter, and their current status within the Red Desert Common name

Latin name

Current status in Red Desert

American badger American goldfinch Barbary sheep Beaver Bighorn sheep Bison Black-footed ferret Bobcat Bonytail Brewer’s sparrow Burrowing owl Bushy-tailed woodrat Caribou Carp Cattle Colorado pikeminnow Columbian mammoth Common raven Cutthroat trout Deer mouse Dog Domestic sheep Dusky shrew Elk European starling Ferruginous hawk

Taxidea taxus Carduelis tristis Ammotragus lervia Castor canadensis Ovis canadensis Bos bison Mustela nigripes Lynx rufus Gila elegans Spizella breweri Athene cunicularia Neotoma cinerea Rangifer tarandus Cyprinus carpio Bos taurus Ptychocheilus lucius Mammuthus columbi Corvus corax Oncorhynchus clarkii Peromyscus maniculatus Canis familiaris Ovis aries Sorex monticolus Cervus elaphus Sturnus vulgaris Buteo regalis

Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; introduction to Wyoming once proposed but never attempted Native; currently occupies Red Desert Native; very small herds currently occupy small pockets of Red Desert Native; currently occupies Red Desert but only in domestic herds Native; currently extirpated from Red Desert Native; currently occupies Red Desert Native; currently extirpated from Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Pleistocene native; extirpated from Red Desert ca. 10,000 bc Exotic; currently occupies Red Desert Exotic; currently occupies Red Desert Native; currently extirpated from Red Desert Pleistocene native; extinct ca. 10,000 bc Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; currently occupies Red Desert in domestic and feral conditions Exotic; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; currently occupies Red Desert, primarily in and near towns Native; currently occupies Red Desert

Vertebrate Wildlife of the Red Desert

145

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 146

Table 5.2 continued

Flannelmouth sucker Golden eagle Golden-mantled ground squirrel Gopher snake Gray vireo Gray wolf Great Basin spadefoot Greater sage-grouse Greater short-horned lizard Grizzly bear Horned lark Horse House mouse House sparrow Humpback chub Kit fox Lake trout Least chipmunk McCown’s longspur Mountain lion Mountain plover Mule deer Muskrat Northern flicker Northern goshawk Northern harrier Northern leopard frog Northern pocket gopher Norway rat Olive-backed pocket mouse Orange-crowned warbler Ord’s kangaroo rat Ornate tree lizard Pigeon Pinyon jay Prairie falcon 146

Catostomus latipinnis Aquila chrysaetos Spermophilus lateralis Pituophis catenifer Vireo vicinior Canis lupus Spea intermontana Centrocercus urophasianus Phrynosoma hernandesi Ursus arctos Eremophila alpestris Equus caballus Mus musculus Passer domesticus Gila cypha Vulpes macrotis Salvelinus namaycush Neotamias minimus Calcarius mccownii Puma concolor Charadrius montanus Odocoileus hemionus Ondatra zibethicus Colaptes auratus Accipiter gentilis Circus cyaneus Rana pipiens Thomomys talpoides Rattus norvegicus Perognathus fasciatus Vermivora celata Dipodomys ordii Urosaurus ornatus Columba livia Gymnorhinus cyanocephalus Falco mexicanus

Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; once extirpated, now occasionally ranges into Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently extirpated from Red Desert Native; currently occupies Red Desert Exotic; native form extinct ca. 10,000 bc; currently occupies Red Desert Exotic; currently occupies Red Desert, primarily in and near towns Exotic; currently occupies Red Desert, primarily in and near towns Native; currently extirpated from Red Desert Native to region; does not occupy the Red Desert Exotic; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies forested landscapes within Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; currently occupies Red Desert, primarily in and near towns Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 147

Table 5.2 continued

Pronghorn Pygmy rabbit Raccoon Rainbow trout Rattlesnake Razorback sucker Red crossbill Red fox Red squirrel Ring-necked pheasant Sage sparrow Sage thrasher Sagebrush lizard Sagebrush vole Sharp-tailed grouse Short-faced bear Silver-haired bat Small-footed myotis Snowshoe hare Spotted bat Striped skunk Swift fox Thirteen-lined ground squirrel Tiger salamander Townsend’s big-eared bat Walleye Western chorus frog Western hognose snake Western terrestrial garter snake White sucker White-tailed jackrabbit White-tailed prairie dog Wyoming ground squirrel Wyoming pocket gopher

Antilocapra americana Brachylagus idahoensis Procyon lotor Oncorhynchus mykiss Crotalus viridis, C. oreganus Xyrauchen texanus Loxia curvirostra Vulpes vulpes Tamiasciurus hudsonicus Phasianus colchicus Amphispiza belli Oreoscoptes montanus Sceloporus graciosus Lemmiscus curtatus Tympanuchus phasianellus Arctodus simus Lasionycteris noctivagans Myotis ciliolabrum Lepus americanus Euderma maculatum Mephitis mephitis Vulpes velox Spermophilus tridecemlineatus Ambystoma tigrinum Corynorhinus townsendii Sander vitreus Pseudacris triseriata Heterodon nasicus Thamnophis elegans Catostomus commersonii Lepus townsendii Cynomys leucurus Spermophilus elegans Thomomys clusius

Native; currently occupies Red Desert Native; currently occupies Red Desert Native to region; likely recently expanded into Red Desert proper Exotic; currently occupies Red Desert Native; currently occupies Red Desert Native; currently extirpated from Red Desert Native; currently occupies forested landscapes within Red Desert Native to region; likely recently expanded into Red Desert proper Native; currently occupies forested landscapes within Red Desert Exotic; currently occupies farmed portions of Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Pleistocene native; extinct ca. 10,000 bc Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies forested landscapes within Red Desert Native; currently occupies Red Desert Native to region; likely recently expanded into Red Desert proper Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Exotic; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native to eastern Red Desert; exotic to western Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert Native; currently occupies Red Desert

Vertebrate Wildlife of the Red Desert

147

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 148

Table 5.3. Examples of widespread vertebrates occupying the Red Desert Class

Species common name

Species Latin name

Amphibian

Tiger salamander Western chorus frog Northern leopard frog

Ambystoma tigrinum Pseudacris triseriata Rana pipiens

Reptile

Greater short-horned lizard Sagebrush lizard Western terrestrial garter snake

Phrynosoma hernandesi Sceloporus graciosus Thamnophis elegans

Mammal

Deer mouse American badger White-tailed jackrabbit

Peromyscus maniculatus Taxidea taxus Lepus townsendii

Bird

Horned lark Golden eagle Common raven

Eremophila alpestris Aquila chrysaetos Corvus corax

Note: These are species that are native to most or all of the Red Desert and also occur throughout large portions of North America.

Table 5.4. Vertebrates that are endemic to the Red Desert Common name Endemic species White-tailed prairie dog Wyoming pocket gopher Wyoming ground squirrel Endemic subspecies Least chipmunk (ssp.) Golden-mantled ground squirrel (ssp.) Thirteen-lined ground squirrel (ssp.) Northern pocket gopher (ssp.) Olive-backed pocket mouse (ssp.) Olive-backed pocket mouse (ssp.) Ord’s kangaroo rat (ssp.) Bushy-tailed woodrat (ssp.)

Latin name

Burrows?

Hibernates?

Cynomys leucurus Thomomys clusius Spermophilus elegans

Yes Yes Yes

Yes No Yes

Neotamias minimus minimus Spermophilus lateralis wortmani Spermophilus tridecemlineatus parvus Thomomys talpoides ocius Perognathus fasciatus callistus Perognathus fasciatus litus Dipodomys ordii priscus Neotoma cinerea cinnamomea

Yes Yes Yes Yes Yes Yes Yes No

Yes* Yes Yes No Yes* Yes* Yes* No

Note: These are species or subspecies with a large proportion of their continental range encompassed by the Red Desert. * Technically “torpor,” a shallower metabolic depression than occurs during true hibernation.

148

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 149

Figure 5.1. White-tailed prairie dog, Cynomys leucurus. The white-tailed prairie dog occupies most intermountain basins in Wyoming and northern Colorado, including virtually all of the Red Desert. It is slightly smaller and paler than its Great Plains cousin, the black-tailed prairie dog (C. ludovicianus). Also, unlike black-tails, white-tails are obligate hibernators. Hibernation lasts approximately from October to May, with occasional arousals during exceptionally warm winter periods. White-tailed prairie dogs dig extensive burrow systems and remain in proximity to burrow entrances at all times. Burrows are clustered into loose colonies, or “towns,” usually in grasslands with only low and sparse shrubs. Although preferentially located on flat and open sites with relatively deep soils, white-tail colonies occasionally form on slopes and often encompass small rock outcrops. Forage includes many species of grasses and forbs, as well as large insects. White-tailed prairie dogs are important prey for several Red Desert carnivores, including badgers, black-footed ferrets, golden eagles, and ferruginous hawks. Furthermore, their burrows provide cover for rattlesnakes, sagebrush lizards, burrowing owls, and several other animals. Sylvatic plague, an introduced disease, has caused large die-offs of white-tailed prairie dogs in recent years. The species is also subject to deliberate shooting and poisoning, often to the point of local extirpation. These three sources of mortality have reduced white-tailed prairie dogs to an estimated 20 to 50 percent of historic population size.

desert, while the Great Basin subspecies inhabits the western side. Rattlesnakes exhibit a similar pattern: the common prairie form occupies the eastern Red Desert, but the unique midget-faded form from the Colorado Plateau edges into the southwestern Red Desert.6 Because its harsh climate has precluded intense land conversions like cultivation and large cities, the Red Desert appears to be emerging as a stronghold for many peripheral and widespread species. A century and a half of cultivation in the Great Plains steadily eliminated core historic populations of prairie birds like mountain plover and McCown’s longspur. This increased the proportional size and conservation value of populations remaining, and in some cases thriving, in the Red Desert. Sage sparrows,

pygmy rabbits, and other species suffering from cultivation and exotic weed invasions in their historic Great Basin cores appear to be similarly prospering in the Red Desert. Trends in population sizes and status rankings indicate that the Red Desert now forms a more important center of occurrence for ferruginous hawks, greater sage-grouse, and other widely distributed species than ever before. Viewed broadly in time as well as space, the ranges of many vertebrates appear to be collapsing on the Red Desert. endemic taxa Three full species and eight subspecies of vertebrates are endemic to the Red Desert. That is, the Red Desert forms a large propor-

Vertebrate Wildlife of the Red Desert

149

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 150

Figure 5.2. Wyoming pocket gopher, Thomomys clusius. Pocket gophers in the genus Thomomys form a complex taxonomic and geographic picture in this region. Wyoming and Colorado support four species in this genus, with distributions that overlap and interdigitate in intricate patterns. The Wyoming pocket gopher is the only mammal species whose entire global range lies completely within Wyoming, and completely within the Red Desert as defined here. Very little is known about Wyoming pocket gophers. Individuals are smaller and paler than other Thomomys and posses fewer chromosomes (2n = 46). Also, relative to the more ubiquitous T. talpoides, T. clusius appears to prefer dry and rocky ridgetops as opposed to deeper and moister soils in valley bottoms. Like all pocket gophers, Wyoming pocket gophers probably spend the majority of their lives underground in long, interconnected burrows of their own digging. They feed primarily on roots, corms, and other underground plant parts and probably occasionally are taken by vertebrate predators. They are not thought to hibernate. Since 1857, at least sixty-eight field-caught Thomomys have been initially identified as Wyoming pocket gophers. However, subsequent taxonomic refinements and reanalyses of these specimens have shown that most belong to other species. Only twenty-one specimens, all captured in the Red Desert of Wyoming, have been confirmed as T. clusius. Pocket gopher specialist C. S. Thaeler Jr. reported twelve individual Thomomys from north-central Colorado with forty-six chromosomes each, raising the possibility that T. clusius occurs there. Cursory field surveys in summer 2005 failed to capture any pocket gophers in the known range of T. clusius. Clearly, much more work needs to be directed toward this cryptic and elusive mammal.

tion of the entire continental range of each of these taxa (Table 5.4; see also Figures 5.1–5.3).7 All are mammals, and rather small in size. Also, each is lighter in color than its relatives in adjacent ecoregions, illustrating “Gloger’s rule,” which states that animals in dry environments are typically paler than those in more humid areas. The functional significance of Gloger’s rule is still debated. Perhaps pale animals are better-camouflaged against light desert soils, or darker animals gain solar heat more quickly to dry themselves in humid environments. Regardless, the regularity with which the rule applies to Red Desert endemics indicates that the area exerts pressures on its wildlife occupants that are significantly and consistently different from pressures exerted by adjacent regions. 150

key adaptations Whether best characterized as peripheral, widespread, or endemic, all vertebrates occupying the Red Desert are well adapted to long, cold, windy winters and short, dry, windy summers. Some species rely on classic adaptations proven successful in other deserts. For example, Ord’s kangaroo rat occupies arid sites across the western United States, including the dune fields of the Red Desert, and is physiologically tuned for extreme water conservation. It can live its entire life without drinking, instead acquiring all necessary water from food digestion, and its urine is highly concentrated to minimize water loss. Other small mammals, such as olive-backed pocket mice, deliberately gather dew from vege-

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 151

Figure 5.3. Wyoming ground squirrel, Spermophilus elegans. The Wyoming ground squirrel exists as three subspecies, each occupying a distinct geographic region: S. e. nevadensis in Idaho, Nevada, and Oregon; S. e. aureus in Montana and Idaho; and S. e. elegans in Wyoming, Colorado, Idaho, and Utah. The third subspecies occupies the entire Red Desert, which clearly forms the core of its range. Rather drab in appearance and general in habitat use, S. elegans pursues a typical ground squirrel life history. Activities are primarily diurnal and center on a burrow system, dug by the ground squirrel itself and providing crucial cover from weather and predators. Burrows are also used for hibernation, which extends from September to March. Wyoming ground squirrels are mostly herbivorous, but diets can contain up to 15 percent animal tissue, including many large invertebrates. Field research also indicates sometimes-heavy predation by Wyoming ground squirrels on eggs and hatchlings of ground-nesting birds, including sage grouse. Wyoming ground squirrels are important prey for many mammalian and avian predators, especially when the vulnerable and abundant young squirrels emerge from burrows in early summer. Wyoming ground squirrel burrows are sufficiently large and ubiquitous to serve as cover for many other vertebrates in the Red Desert. Similar to white-tailed prairie dogs in this regard, the Wyoming ground squirrel can be seen as a keystone species in the Red Desert, providing underground refuges in an otherwise harsh and exposed environment.

tation in their fur, which they later lick to ingest water.8 Perhaps the most impressive set of adaptations is found in the Great Basin spadefoot, which is somewhat of a living oxymoron: a desert-loving amphibian. This toad relative usually avoids permanent ponds and streams and is commonly found many miles from reliable water. Spadefoots use hardened nails on their hind feet to burrow several feet into soft soil (occasionally they descend mammal burrows instead), where they secrete a waterproof cocoon and enter metabolic torpor, slowly dehydrating and living off stored fats. There they persist through winter and through long summer dry spells. Warm and wet conditions, perhaps following a brief summer thunderstorm, trigger emergence and a flurry of rehydration, feeding, and mating. Eggs hatch

in two to three days, and tadpoles metamorphose in as little as four weeks, one of the fastest egg-to-adult transformations known. Surface activity continues until the water body — maybe only a rain puddle—begins to dry, sending all spadefoots back into the soil.9 Ten months may pass before the next opportunity to frantically emerge, breed, and reburrow. Ten of the eleven endemic mammals listed in Table 5.4 dig burrows and spend much of their time underground. The exception, the bushy-tailed woodrat, builds a burrowlike structure aboveground: a grass-lined nest, inside a large pile of debris, inside a cave or rock crevice. The Red Desert clearly favors animals that place physical barriers between themselves and the weather. Relative to the ground surface, burrows are warmer in winter,

Vertebrate Wildlife of the Red Desert

151

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 152

cooler in summer, and more humid year-round, owing to stable soil temperatures and moisture content as well as the absence of wind. These effects are accentuated when occupants plug burrow entrances with dirt, as they often do. Many other Red Desert animals, in addition to the endemic mammals, dig burrows, most notably American badgers, pygmy rabbits, sagebrush voles, and even sagebrush lizards. And still more animals rely on the burrows of these primary excavators: burrowing owls, rattlesnakes, and Great Basin spadefoots are well-known “secondary” burrow users. Thus, primary excavators can be seen as keystone species in this area; their actions significantly alter landscape architecture to the benefit of many other species. By adding burrows to the Red Desert, they provide a protected vertical dimension to an otherwise exposed landscape. And burrowing also contributes substantially to soil aeration, water infiltration, and nutrient mixing, which ultimately affect the diversity and productivity of vegetation.10 Eight of the eleven endemics hibernate through the long Red Desert winter, as do many other species—some bats, for example, such as the small-footed myotis and Townsend’s big-eared bat.11 But others, such as the spotted bat and the silver-haired bat, contend with winter as do many birds that breed in the Red Desert: they migrate. The Red Desert simply does not produce enough food in winter for many small vertebrates, especially those that are highly insectivorous. Maintaining a consistent body temperature would be impossible for small endotherms like bats and birds, which cannot accumulate much fat due to the weight limits of flight and have large surface areas that quickly radiate body heat. In contrast, many large mammals are challenged more by overheating than by cold, because their voluminous body cavities produce more heat than their proportionally small surface areas can easily radiate. With enough quality forage, almost any large mammal can endure extreme cold. For example, well-fed large African mammals survive fine in relatively frigid North American zoos. Thus in winter some pronghorn, mule deer, and elk mi152

grate into the Red Desert, where they can fuel their large bodies with abundant exposed forage. This is much less expensive, energetically, than remaining in the foothills and mountains to travel and dig for forage through deep snow. Large well-fed mammals can also consume snow for water without incurring much of an energetic cost. Not so for birds and smaller mammals. They must invest relatively large amounts of metabolic energy to raise consumed snow to body temperature— yet another reason why most small endotherms avoid Red Desert winters by hibernating or migrating. Critical Landscape Features In any landscape there are a few physical or biological features that substantially increase local biological diversity by providing cover, water, and/or food in a manner that is only poorly duplicated, if at all, by other features. Mammal burrows are a good example of such a feature in the Red Desert: as discussed above, burrows provide critical cover for a host of secondary users. ponds and flowing streams Open water is rare in any desert, and there is little argument that the biota of the Red Desert would be impoverished without its precious few permanent ponds and streams. Fish would be entirely absent, along with semiaquatic mammals (e.g., beaver, muskrat) and most amphibians. Permanent ponds and streams allow waterfowl and shorebirds to breed in the Red Desert and also to stop over during fall migration. Without such waters these birds would appear only briefly, on seasonal ponds, during spring migration.12 Even some purely terrestrial vertebrates would be hard-pressed to occupy the Red Desert without permanent water. Virtually all bats in the region feed on insect swarms produced by wetlands; bats commonly travel several miles each day between upland roosts and wetland foraging sites. Without wetlands that last into late summer most bats would not be able to build up the

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 153

nutritional reserves necessary for their impending hibernation or migration. Cottonwoods, willows, sedges, and tall grasses like Great Basin wild rye occur almost exclusively near ponds, seeps, and streams, where their roots can easily tap subsurface water. Riparian vegetation remains green late into summer when many upland plants have already cured.13 Songbirds such as orangecrowned warblers and American goldfinches may not occupy the Red Desert at all, or perhaps would stop only fleetingly during migration, in the absence of riparian thickets.14 Perennial ponds and streams are most important in late summer, when seasonal ponds and streams are dry, vegetation is cured, temperatures and evaporation rates are high, and snow has yet to fall. This is when larger species such as pronghorn, mule deer, and elk (and livestock) concentrate around remaining water sources. Distributions of smaller species, too, like white-tailed jackrabbits and greater sage-grouse, increasingly shift toward open water. Any water remaining in the Red Desert from midAugust to mid-October is the setting for classic wildlife scenes, with herds and flocks nervously approaching muddied pools, and animals jockeying for shoreline position, all under the watch of predators that have staked out the last water holes. mama sage Sagebrush is absolutely central to Red Desert wildlife, as evidenced by the five vertebrates (and at least as many invertebrates) having “sagebrush” or “sage” in their common names: sagebrush lizard, sagebrush vole, greater sage-grouse, sage thrasher, and sage sparrow. In general, sagebrush provides wildlife with many of the same benefits of other shrub species—leaves and seeds for forage, a woody skeleton for cover, and stiff branches for perches. All shrubs, including sagebrush, act as snow fences in winter, trapping snow in leeward drifts that melt slowly and allow water to percolate deep into the soil in spring. Much of this snow would otherwise blow away, sublimate, or melt too quickly and evaporate.15

But sagebrush is not just another shrub, either aesthetically or functionally. It is the only shrub in the region that retains metabolically active leaves through the winter.16 Greater sage-grouse and pygmy rabbits eat essentially nothing but these leaves in winter, the latter sometimes climbing up several feet into the shrubs to feed. Ungulates, too, rely heavily on sagebrush as winter forage.17 The winter leaves are nutritious and grow on stiff twigs that protrude above the snow; thus they can be eaten without the searching and digging required by other winter forage like desiccated grasses and forbs. Winter leaves also make for tighter shrub canopies, increasing the effectiveness of sagebrush stands as winter cover. Exceptionally tall subspecies of sagebrush, like basin big sagebrush (with some shrubs approaching ten feet in height), provide treelike cover all year. Elk herds in the Jack Morrow Hills regularly use stands of basin big sagebrush for shade in hot weather and as cover from disturbances year-round.18 mountains, buttes, and rims True mountains such as the Crooks-Green-Ferris chain along the Sweetwater River, and the Little-Pine – Cold Spring mountain country near the junction of Colorado, Wyoming, and Utah, support boreal environments and boreal forms of wildlife, thereby increasing the biological diversity of the Red Desert as a whole. Snowshoe hares and red squirrels, for example, occupy the Green and Ferris mountains alongside red crossbills and northern goshawks. Populations of boreal mammals on these mountain islands are isolated, at least partially, from populations on the “mainland” Rocky Mountains, raising the possibility that they have developed unique genetic signatures.19 This issue has received essentially no research attention to date. Large mountains, small mountains, buttes, and canyons provide cooler summer range for elk and mule deer, as well as roosts and hibernacula for several bats. These features are also important to bighorn sheep, which rely almost completely on steep and rugged terrain to escape predators. Historical observations of

Vertebrate Wildlife of the Red Desert

153

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 154

bighorn sheep in the Red Desert were always recorded on or near such features,20 and any future recovery of bighorn sheep in the area will necessarily center on such features. All canyons and highlands, including small ridges and rims, interact with the wind in important ways. In addition to providing windbreaks for animals in all seasons, most of them capture blowing snow in drifts that persist into the spring, melting slowly and providing moist sites for a variety of plants and animals. Drifts form in hollows and on leeward slopes; in contrast, ridgetops commonly stay windblown and snow-free. Here, sagebrush and other forage plants remain accessible through heavy winters, which is extremely important to pronghorn and mule deer. Rims and ridges play subtle but important roles for smaller species too. For example, mountain plovers commonly nest on sharp rims and ridgetops in the Red Desert. Wind scour causes these sites to support very little vegetation (mountain plovers prefer to nest in barren areas). Perhaps most importantly, mountain plovers can easily see aerial predators approaching such locations. Prairie falcons, northern harriers, and other raptors must skyline themselves when approaching highland crests.21 trees Trees are rare and scattered in the Red Desert, and associated to a large degree with major topographic features. Large mountains support forests of boreal conifers; smaller highlands support isolated stands of montane conifers and aspen, often in sites that accumulate snowdrifts. Juniper woodlands occur on some ridges and rims — juniper berries are important sources of food and water for many rodents and birds, especially in winter.22 Pinyon pine edges into the southwestern Red Desert, and its large and nutritious seeds help support several birds, such as pinyon jays, that either do not occur or occur at much lower densities elsewhere in the region.23 Small stands of trees also occur in moist situations separate from mountains and buttes. Cottonwood stands form at intervals 154

along larger permanent streams; smaller streams and seeps sometimes support isolated stands of aspen, along with mesic shrubs like chokecherry and willows that can approach trees in height and structure. Although rare and patchy in distribution, trees can be discussed in the same terms used earlier for animal burrows in the Red Desert—they provide a vertical dimension to an otherwise two-dimensional landscape, and support primary excavators, such as northern flickers and other woodpeckers, that produce protected cavities for themselves and a variety of secondary users. Invertebrates The invertebrates of the Red Desert are poorly known, and there is a clear need for more research directed toward them.24 Beyond the many basic ecological functions they perform, like pollination and decomposition, it is almost certain that the Red Desert supports invertebrate species that occur nowhere else. The Killpecker Dunes, for example, immediately suggest a site where unique insects may have evolved. Several vertebrates depend heavily on invertebrates as food. Shrews, bats, and many birds are purely insectivorous, as are some amphibians and reptiles. Other vertebrates are partially insectivorous, including some (like chipmunks and pocket mice) typically described as herbivores, and others (like foxes, bobcats, and most raptors) assumed to prey almost exclusively on other vertebrates. The physical conditions of the Red Desert probably place special value on invertebrates as prey during particular times of the year. Vertebrates generally require much protein in spring and early summer. Birds, for example, are producing protein-rich eggs, and hatchlings require much protein for rapid growth. Pregnant and lactating mammals, and their quickly growing offspring, are similarly protein-hungry. But plant protein is rare at this time of year, as new vegetation grows slowly during the cold Red Desert spring. Emerging invertebrates, collectively, may represent a crit-

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 155

ical protein source at this time for many otherwise herbivorous vertebrates. It is well established that greater sage-grouse chicks require much insect prey in the first two weeks after hatching or they suffer high mortality, and also that such mortality is a significant limit to population growth.25 This scenario probably extends well beyond greater sage-grouse. The reproductive success of many other vertebrates may be constrained by protein availability in the spring, and, in the Red Desert, springtime protein equals insect protein. In late summer, the driest part of the year, invertebrates may be critical sources of water. Many vertebrates probably target insects (especially large, fatty insects like crickets, darkling beetles, and grasshoppers) for their high water content at this time. Descriptions of the diets of many vertebrates include phrases like “a few insects eaten” and “insects taken occasionally.” Such casual phrases may discount the importance of invertebrate prey. Although they may form only a small proportion of the annual diet, invertebrates might still bridge critical nutritional gaps during certain seasons. Prehistory of Red Desert Wildlife The deep past of the Red Desert, like that of any place, is populated by bizarre animals in fantastic settings: allosaurs stalking through Jurassic swamps, rhinolike uintatheres grazing near giant Eocene lakes, and Pleistocene megamammals, such as Columbian mammoths and giant short-faced bears, enduring an even colder, drier, and windier Red Desert than we know today.26 The current fauna of the Red Desert resolved about 10,000 – 8,000 bp,27 when the megamammals became extinct and a small but rather effective new predator—Stone Age humans— entered the desert for the first time.28 Relatively minor climatic changes have occurred over the past 10,000 years,29 all of which affected the distribution and abundance of Red Desert biota. But western North America was rather

open during most of this period, in the sense that plant and animal populations could move rather freely across large areas with few hard barriers to migration or distributional shifts. The rise of cities, highways, and other artifacts of technology-dependent humans have progressively closed the region. In the short term this threatens seasonal migrations. In the long term it endangers distributional shifts required to survive climatic changes. And there is no legitimate doubt that the climate is changing again, forcing wildlife across the globe to move in response to the unprecedented warming produced by the same technology-dependent humans.30 Maintaining long-distance movements of wildlife in an increasingly human-dominated landscape is a core issue in current debates over the Red Desert. Early History: 500–200 bp The period 500 –200 bp was a time of accelerating environmental change in North America, as European colonization moved across the continent. But by 1800 the Red Desert and its wildlife had felt only distant and secondary effects of the expansion. The Spanish introduced the horse to North America in the mid-1500s (or was it a reintroduction, since native horses roamed the Red Desert alongside other soon-to-be-extinct megafauna at the end of the Pleistocene?),31 and wild horses soon occupied the Red Desert.32 From 1600 to 1800 many indigenous groups adopted the horse as a central technology, which allowed for longer hunting and trading forays, easier travel across dry uplands, and perhaps most important, the ability to follow bison herds year-round.33 The new pressures of wild horses and horsemounted humans undoubtedly affected Red Desert wildlife, especially big game, in complex ways. At this same time indigenous people, and their effects on wildlife, may have been affected by the first waves of European diseases passing between Native American groups far in advance of Europeans themselves.

Vertebrate Wildlife of the Red Desert

155

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 156

European Contact: 200–100 bp Whereas the responses of Red Desert wildlife to environmental changes before 1800 are somewhat speculative, there is no question that wildlife populations here suffered at the hands of people during the 1800s. Stream systems in the outer Red Desert were heavily trapped for beaver in the early 1800s. References to the Green River and its tributaries are omnipresent in trapper accounts of the era. Beaver damming of streams produces many keystone effects: storage of water, sediment, and nutrients; higher late-season water flows; habitat creation for certain fish, amphibians, waterfowl, and shorebirds; more diverse and productive riparian vegetation.34 It is presumed that these effects were largely suppressed by the mid-nineteenth century as beaver were trapped to the point of extirpation in many locales.35 Water tables likely dropped as remaining beaver dams failed, reducing the coverage of riparian vegetation and causing some stream downcutting, all to the detriment of native wildlife.36 Streamside vegetation and wildlife continued to decline under heavy and persistent grazing by emigrant livestock. Riparian shrubs and grasses can survive intense grazing by large mammals as long as it occurs only once every few years, as bison, elk, and other wildlife are thought to have grazed Red Desert stream corridors. But almost all stream segments near the emigrant trails would have been grazed yearly, and intensively, by livestock beginning in the 1840s.37 Willow, currant, Great Basin wild rye, and other riparian plants would have declined in abundance and vigor under such pressure, leading to subsequent declines in the native vertebrates that depend on such plants. The effects of Euro-American expansion on wildlife in the Red Desert were probably rather localized until about 1870. Key technological infrastructure such as the transcontinental telegraph (completed 1861) and the Union Pacific railroad (completed 1869), which bisected the Red Desert, greatly increased the mo-

156

tivation and ability of people to convert the natural capital of the region into market capital. At this point, immense herds of sheep and cattle began to be driven into every corner of the Red Desert. Riparian meadows throughout the region received the heavy year-round grazing formerly restricted to the emigrant trails. In the late 1880s many were converted to hay meadows, and some even to small crop farms. The responses of certain native vertebrates were predictable. For example, Columbian sharp-tailed grouse, a small subspecies of sharp-tailed grouse, historically occupied stream corridors throughout southwestern Wyoming and northwestern Colorado. The subspecies is now quite rare, and published assessments consistently point to livestock-induced changes in riparian vegetation as a main cause of the decline.38 An isolated population of Columbian sharp-tailed grouse currently occupies the foothills along the southeastern fringe of the Red Desert, in what ecologists generally refer to as “mountain shrub” vegetation. However, there is some speculation that this term may be historically inaccurate. This productive mix of mesic shrubs and tall grasses may have once extended far into sagebrush-dominated basins along stream corridors, but could not endure there in the face of persistent livestock grazing. It is now restricted to the foothills, where higher precipitation and steeper slopes make it more resilient to grazing pressure. Livestock began to pressure vegetation and wildlife in the uplands as well. Millions of cattle and sheep removed huge amounts of plant biomass that would have otherwise served as forage and cover for wildlife.39 Excessive trampling of soil, vegetation, and even bird nests by livestock were commonly noted by early stockmen.40 More subtle changes, too, were observed by those close to the land. Dr. David Love, a noted Wyoming geologist, has pointed out how early trail herds dramatically increased the number and type of selenium-accumulating plants in the Red Desert. By concentrating naturally occurring selenium and releasing it to the topsoil, these plants eventually raise selenium

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 157

levels in much of the surrounding vegetation. This lowered forage quality, sometimes to the point of toxicity, for livestock and native wildlife.41 The best-known wildlife decline of the late nineteenth century in the Red Desert was the near extinction of big game, especially bison. The classic story centers on overkill: huge game herds fell to market hunters, hide hunters, and “recreational” shooters during the last half of the century.42 But exotic diseases introduced from expanding herds of livestock share some of the blame. As early as 1825, bison were dying in large numbers from apparent livestock afflictions in Nebraska.43 This was followed by die-offs further west, including, in 1858, an epidemic of “bloody murrain” (an old term for anthrax) decimating bison, deer, and pronghorn from Fort Laramie to Fort Bridger right through the heart of the Red Desert.44 Tuberculosis, brucellosis, and “tick fever” (a catchall term that likely encompassed several livestock diseases) are variously mentioned as moving from livestock to native game during this period.45 Scab (scabies) from domestic sheep spread through Wyoming bighorn sheep at a frightening rate following its first report in about 1880;46 outbreaks in Colorado may have started even earlier.47 Riparian degradation, forage competition from livestock, overkill, disease, and the disastrous winter of 1886 –1887 brought all big-game herds in the Red Desert to their historic low points between about 1885 and 1908. In 1906 the Wyoming state game warden estimated that only 2,500 pronghorn occupied the entire state, with most in the Red Desert.48 It is generally agreed that the last indigenous, free-ranging herd of bison outside of Yellowstone National Park occupied the Red Desert and was extirpated around 1889.49 Later reports of Red Desert bison are almost certainly of escapees from various domestic herds. As late as 1967 a small group of bison roamed the core Red Desert, under speculation that they were remnants of the original wild herds. But David Dary quotes A. W. Reynolds of the Wyoming state legislature, who suggested in 1969 that they were fugitives: “A few years back, a herd of buffalo, escaped from a private herd on the Sweet-

water river north of Wamsutter, started developing in the northeast corner of the Red Desert, and were reported by the game and fish boys to be, variously, from eighteen to seventy-five head. No recent reports, and the rumor is that the uranium miners have enjoyed buffalo hump and tongue.”50 Human-caused declines in big game were accompanied by declines in native predators. In the Red Desert, grizzly bears probably never ranged far from the riparian environments treasured by stockmen. They were rapidly eliminated through habitat degradation, prey reduction, and bounty-enhanced shooting and poisoning. Gray wolves were also deliberately targeted as stock killers, especially as their native prey were increasingly replaced by cattle, sheep, and horses.51 Even raptors were placed under bounties to reduce predation on lambs. Smaller mammals and birds suffered as well, mostly from range overgrazing, but also as secondary victims of poisons set for larger species. There were a few additions to the Red Desert biota during this period. Russian thistle likely reached the Red Desert by 1880, just ahead of other noxious weeds, such as cheatgrass and halogeton, that are now firmly established.52 In the 1880s the State of Wyoming began aggressively stocking many exotic fish (including species later targeted for removal, such as common carp) in waters across the state, including the North Platte and Green river systems.53 House mice and Norway rats likely accompanied people into towns, as they had been doing for millennia. Horses were not new, but herds tended by Native Americans and stockmen swelled and sent more escapees into feral bands. Feral dogs, also not new, probably saw a similar boost in numbers. Twentieth Century: The Start of Recovery The decline of the western big-game herds helped develop a conservation ethic at local and national scales. Much of this was motivated by moral concerns, but some were economic: by 1900, big-game hunting and wildlife viewing were recognized as major economic forces, as was more enlightened management of soil,

Vertebrate Wildlife of the Red Desert

157

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 158

water, and other natural resources. States established formal wildlife agencies and laws protecting wildlife and habitat, and scientific techniques of wildlife management were increasingly developed and applied. The first wildlife regulations enacted in Wyoming and Colorado would have been unthinkable only a few years earlier, including basic ones now taken for granted like designated hunting seasons and bag limits. Some of the more aggressive early management actions would be politically difficult even today; for example, statewide bans on beaver trapping, multiple-year closed seasons on greater sage-grouse, and publicly funded land purchases for game preserves.54 But in general these actions worked. Big-game herds rebounded across the west, with most species forming huntable and actively managed herds in much of their historic ranges by 1915.55 This was the case for Red Desert pronghorn and mule deer, whose populations thrived during the twentieth century. Indeed, the Red Desert now supports the world’s largest pronghorn herd. But bison were effectively gone, and bighorn sheep and elk were still far below historic levels. Curiously, the Wyoming state legislature passed a law protecting “wild” bison in 1899,56 ten years after the last wild herd anywhere (excluding Yellowstone National Park) was extirpated in the Red Desert. This may have been an attempt to reestablish free-ranging bison from occasional domestic escapees. In his 1916 annual report the Wyoming state game warden raised the issue of “re-stocking” Wyoming with bison, which probably seemed like a natural follow-up to the successful recovery of other game species.57 The proposal clearly failed. In the mid-1960s a Wyoming legislator twice proposed bills to protect the fugitive bison then roaming the Red Desert as game animals under the management of the Wyoming Game and Fish Department. But these failed too, largely because stockmen feared transmission of brucellosis from bison to cattle.58 The historical irony that brucellosis may be one of the livestock-introduced diseases partially responsible for the near extinction of the bison in the first place 158

is not lost on some participants in the current debate over bison, elk, and cattle management in western Wyoming. Elk herds in the Yellowstone and Jackson Hole regions persisted through the late nineteenth century in relatively good condition and became sources for restocking elk across the west. Several early conservationists recognized the important link between these elk herds and the Red Desert. Archaeological evidence and firsthand observations from the late 1800s make it clear that huge numbers of elk (and pronghorn and mule deer) historically migrated long distances from summer range in the mountains of western Wyoming to winter range in the Red Desert.59 Many mule deer and pronghorn still undertake such migrations.60 In fact, herds that winter in the western Red Desert migrate farther than almost all other ungulate herds in North America, with the pronghorn migration between Grand Teton National Park and the Red Desert exceeded in length only by migrations of barren ground caribou in northern Canada.61 These movements are particularly impressive given the number of longestablished obstacles in the region such as fences, highways, and cities, plus the newer challenges of oil and gas extraction fields and rural subdivisions. But elk are more sensitive to such obstacles, and elk migrations to the Red Desert were disrupted by human developments in the Green River valley by 1900. Occasional stragglers attempted the journey as late as 1921.62 Cut off from the desert, large numbers of elk concentrated on smaller and less-suitable winter ranges at higher elevations, such as Jackson Hole itself, which started a cascade of management problems: unnatural levels of winter mortality, conflicts with stock growers on deeded mountain ranches, artificial feeding and feed grounds for elk, and high incidences of brucellosis (and, perhaps soon, chronic wasting disease).63 Viewed from a historical perspective, many of the controversies that dominate game management in northwest Wyoming today stem directly from blocking traditional elk migrations to the Red Desert.

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 159

Recognition of the importance of these migrations led to several large-scale conservation proposals. In an 1898 issue of Recreation magazine Dr. Frank Dunham proposed the Red Desert as a national game preserve, linked via protected migration corridors to Yellowstone National Park. The idea occasionally resurfaced in various forms; the Wyoming state game warden advocated a formal Red Desert game range in 1907, and the Wyoming state legislature protected some land for game in the area in 1909.64 In 1935, Wyoming governor Leslie Miller proposed the Jack Morrow Hills portion of the Red Desert as a national park, and in 1961 the U.S. Congress considered the area for national monument status. It is assumed that these proposals were motivated at least in part by a desire to protect game herds.65 A small elk herd may have occupied Steamboat Mountain in the Red Desert into the 1930s, but it was likely nonmigratory. In an attempt to reestablish the traditional migrations, the Wyoming Game and Fish Department began moving elk from the National Elk Refuge to the western Red Desert in 1947, ultimately releasing several hundred animals over the next twenty-one years. At first some of these elk began regular, seasonal movements between the mountains and desert, but by 1952 the migrations had stopped and the elk remained year-round in the low country.66 Red Desert elk have thrived over the past twenty years, paralleling trends in most western elk herds, and there are anecdotal accounts of a few individuals making seasonal treks to the mountains. This herd is prized by many people and figures centrally in debates over management of the area. Desert elk, with brightantlered bulls (antlers rubbed on sand and sage remain grayish white, in contrast to antlers rubbed on pitch-filled conifers in the high country) bugling amidst herds of pronghorn and freeranging horses, have a primal attractiveness that many are unwilling to sacrifice. And the primeval nature of such scenes has only increased in recent years as gray wolves, dispersing south from reestablished populations in northwest Wyoming, now make regular winter forays into the northwestern Red Desert.

New and Continuing Challenges The post–World War II boom in technology and human population has challenged Red Desert wildlife in numerous ways, some of them new, others merely continuations of older processes. Exotic plants and animals have clearly increased in the area. Cheatgrass, an aggressive Eurasian annual that has overrun huge swaths of the Great Basin, is thought by some to be poised to take over the Red Desert. However, although cheatgrass clearly occupies portions of the desert, it does not appear to be expanding rapidly. It may be poorly adapted to the unique combination of aridity and cold found here.67 But the climate is changing, and annual grasses can rapidly evolve to new environments. The future of cheatgrass in the Red Desert is rather an open question. Fish communities in the outer Red Desert have been substantially changed by deliberate and inadvertent introductions of game, bait, and “trash” species, including some originally from Europe and Asia. Large reservoirs such as Seminoe and Pathfinder on the North Platte River, and Flaming Gorge on the Green River, support lacustrine species (e.g., walleye, lake trout) that otherwise could not persist in the area.68 Many exotic fish prey on and compete with native species; additionally, some hybridize with natives, such as exotic rainbow trout crossing with Colorado River cutthroat trout, and exotic white suckers crossing with flannelmouth suckers.69 Furthermore, development of sport fisheries in the twentieth century commonly involved deliberate fish poisoning. The construction of Flaming Gorge Reservoir in 1962 was attended by the poisoning of 445 miles of the Green River with an estimated 21,000 gallons of rotenone to remove nongame fish and prepare for the stocking of game species.70 This poisoning, along with the movement-blocking effects of the dam and reservoir, eliminated much range for four fish species whose remaining populations downstream of Flaming Gorge are now under Endangered Species Act protection: Colorado pikeminnow, razorback sucker, bonytail, and humpback chub.71

Vertebrate Wildlife of the Red Desert

159

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 160

Fish poisoning still occurs around the perimeter of the Red Desert, but now it is often used to remove exotic trout (some deliberately stocked in previous decades) to expand populations of, and reduce hybridization with, Colorado River cutthroat trout. Similarly, the recently completed High Savery Reservoir about thirty miles south of Rawlins, Wyoming, is first and foremost a water storage structure but is also being managed to produce Colorado River cutthroat trout for reintroduction elsewhere. European starlings, house sparrows, and pigeons now occupy the Red Desert, primarily in the vicinity of human habitations. Ring-necked pheasants have been stocked heavily throughout Colorado and Wyoming for at least eighty years, and some occasionally edge into milder portions of the Red Desert. Foreign sport fish and game birds are likely to remain in the area and benefit from active management. But in general the midtwentieth-century enthusiasm for stocking exotic sport animals (even Barbary sheep from North Africa were briefly considered for wild release in Wyoming in the 1960s72) appears to have been replaced by a desire to recover and conserve native species. Some species that are native to western North America, but historically absent from the Red Desert, have moved into the area over the past fifty years. Anecdotal observations suggest that raccoons and striped skunks were rare in the Red Desert until the last few decades; perhaps the first towns and farms in the region served as beachheads for initial population establishment and now support enough reproduction to supply dispersing individuals out into undisturbed desert. Red foxes were also rare in the Red Desert until recently. Historically, native red foxes probably ranged only infrequently into the desert from their preferred haunts in adjacent mountains. But by 1960 a lowland-adapted red fox (probably originally imported into the eastern United States from Europe) expanded from the Great Plains into Wyoming and Colorado and is now commonly seen in intermountain basins.73 Raccoons, skunks, and foxes are notorious predators of eggs and hatchlings, and their effects on greater sage-grouse and other native Red Desert birds have not yet been thoroughly studied. 160

Livestock are still essentially ubiquitous in the Red Desert, but the intensity of livestock use is arguably less now than before 1934, when the Taylor Grazing Act ended free-for-all, open-range grazing. Cattle and sheep still remove significant amounts of forage and often concentrate around water sources to the point of degrading water quality and trampling stream banks. Although practically and politically difficult, management of livestock grazing in such a way as to prevent damage to native vegetation and wildlife is possible. It is more difficult, however, to manage the many feral horses now ranging the desert. The impact of horses on vegetation and water sources is much the same as that of cattle, but unlike cattle, feral horses are not the concern of private owners, and unlike big game, they are not managed by state wildlife agencies. Currently, feral horses in the Red Desert are controlled via periodic roundups and auctions conducted by the U.S. Department of the Interior (USDI) Bureau of Land Management. There is a rather powerful lobby, backed by federal legislation (the Wild Free-Roaming Horses and Burros Act of 1971), that advocates expansion of feral horse herds in the western United States.74 Several diseases still challenge Red Desert wildlife. Sylvatic plague, originally from Asia, and canine distemper are major killers of black-footed ferrets. White-tailed prairie dogs, the ferret’s main prey in the Red Desert, also suffer from plague epidemics, which greatly compound the damage populations suffer from widespread shooting and poisoning. Black-footed ferrets were likely gone from the desert by 1980 and were probably globally extinct, save eighteen captive individuals, by 1985.75 There are now eight reintroduced wild populations of ferrets in North America, two of which occur just outside of the Red Desert. All are intensively managed, with much attention focused on detecting and mitigating the effects of disease on ferrets and their prairie dog prey. Bighorn sheep in the Red Desert never recovered from their late-1800s crash, mostly due to continued contact with diseases from domestic livestock. Written accounts of fur trappers and

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 161

early emigrants suggest that bighorn sheep were abundant and widespread in the Red Desert prior to 1870.76 But populations quickly succumbed to a variety of livestock diseases, including pasteurella, scabies, and lungworm,77 and wildlife managers today are resigned to the fact that bighorn sheep will thrive only where there is minimal contact with cattle and, especially, domestic sheep. Again, livestock can be managed to benefit wildlife, including the creation of “safe zones“ for bighorn sheep populations, but it requires a willingness to overcome certain practical and political obstacles. Many of the challenges facing Red Desert wildlife today—indeed, many of the challenges facing wildlife anywhere in the twenty-first century—share a common theme: roads. Roads generate a well-known suite of ecological impacts in almost all environments, and the Red Desert is now permeated by roads of all varieties.78 The most obvious effects —vehicle pollutants, road kill, replacement of native vegetation by road surfaces—are less important than more subtle ones. For example, roadsides are ideal strongholds for exotic weeds. Vehicles continually import weed seeds from distant sites into the disturbed and relatively wellwatered soils next to roads. Halogeton (and in some places, cheatgrass) now forms obvious “stripes” bordering many Red Desert roads. It is more difficult to directly observe the profound changes in animal behavior induced by roads. Vehicle movement, mechanical noise, dust, shooting (including poaching), domestic dogs, offroad traffic, and other stimuli eventually cause many vertebrates to avoid roads by considerable distances. Densities of some sagebrush-dependent songbirds, such as Brewer’s sparrow and sage sparrow, are 50 percent lower near roads than in areas more than a hundred yards distant from roads in southwest Wyoming.79 In the same area, greater sage-grouse hens that breed near roads move farther before attempting to nest, and end up nesting less frequently than hens that breed in unroaded areas.80 Interestingly, some large vertebrates can habituate to certain road types, as evidenced by road-tame carnivores and ungulates

in national parks and large herds of pronghorn and mule deer occasionally congregating near interstate highways. But the secondary and tertiary roads now penetrating the Red Desert produce stimuli, like shooting and all-terrain vehicles, that do not typically emanate from park roads or interstates. Recent studies show that elk in the Red Desert regularly avoid areas within 1.7 miles of roads and 1.2 miles of active oil and gas wells,81 which are linked by roads and produce many of the same stimuli as roads. Similarly, mule deer in southwest Wyoming avoid areas within 2.3 miles of active oil and gas wells.82 These large buffer distances are understandable in such open country. In general, large mammals can screen themselves from disturbances with tall vegetation, sheer distance, or steep topography. But tall vegetation is rare in the Red Desert, forcing more reliance on distance and topography. As the growing road network reduces the amount of land distant from disturbances, large mammals must increasingly rely on topographic features such as buttes, rims, and canyons for security. Now even these features are becoming roaded, as rough topography attracts some users of all-terrain vehicles. Road impacts on large carnivores, such as grizzly bears, gray wolves, and mountain lions, have not been directly investigated in the Red Desert, but the general story is clear. Multiple studies from across western North America indicate strong avoidance of roads by large carnivores and indicate further that the primary limit on populations of these species is human-caused mortality, which in turn is a function of human density, which in turn is a function of road density.83 Interstate 80 and its associated frontage roads, railroads, wildlife-proof fences, and other artificial structures pose a major barrier to wildlife movement.84 In fact, from the perspective of many reptiles and mammals, Interstate 80 has effectively created two separate Red Deserts — one north of the highway, and one south of it. Some individuals traverse the highway at stream crossings, and others use the few small and widely spaced underpasses designed for wildlife passage. But for most practical purposes the

Vertebrate Wildlife of the Red Desert

161

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 162

majority of reptile and mammal populations separated by Interstate 80 are isolated from one another. Roads and other movement barriers affect large mammals the most during periods of heavy snow. Biologists often document large numbers of pronghorn and mule deer dying at tight roadside fences during and after heavy snowfalls in the Red Desert. More than 10,000 pronghorn drifting south from the 1983–1984 blizzards were stopped by highway fences at Rock Springs, Wyoming, triggering massive attempts to move and feed stranded animals. Ultimately, several hundred died.85 During the same winter a famous incident played out at Red Rim southwest of Rawlins, Wyoming. About 1,600 pronghorn were stopped by a new wildlife-proof fence enclosing winter range on a mosaic of private and federal lands. Media coverage, intense public reaction, and negotiations involving the governor of Wyoming eventually led to the removal of several sections of fence, but spring surveys indicated that around 1,000 animals may have died. This incident forced more emphasis on proper design of rangeland fences that restrain livestock but allow pronghorn passage.86 Some artificial features may attract wildlife during heavy snowfalls. For example, pronghorn sometimes congregate on railroad tracks kept free of snow by wind and passing trains (often the animals access the tracks by walking over drift-covered fences). In 1949 more than 800 pronghorn were killed by trains under such circumstances in southwest Wyoming.87 The Wyoming Game and Fish Department reported 41 pronghorn killed by a single train near Granger, Wyoming, on November 17, 2003, under relatively mild snow conditions. Roads sit at the heart of the most contentious activity in the Red Desert today: energy development. Oil, natural gas, coal, and uranium have been extracted from the Red Desert for decades, but the scales of current and planned projects, especially for oil and gas, dwarf any past activity. When these projects are being considered, it is vital to remember that their impact on wildlife is not measured by their direct footprints—that is, the amount of ground that is actually disturbed—but rather by their larger dis162

turbance footprints that account for species’ avoidance of wells, roads, and associated features. Analyses of oil and gas fields in the upper Green River valley indicate that direct footprints covering less than 5 percent of a field can produce disturbance footprints covering over 95 percent. Importantly, these figures are based on buffer distances of only 0.25 mile,88 which is considerably less than the 1- to 2-mile avoidance buffers measured for elk and mule deer. And road effects clearly extend through time as well as space; roads, and their attendant disturbances, persist long after mineral extraction ends. Aerial photographs of coal-bed methane fields reveal shockingly dense road networks. In fact, in contrast to conventional petroleum fields, where the amount of native vegetation directly replaced by road surface is of only moderate concern, coal-bed methane fields encompass road networks that are so dense as to represent a significant conversion in land cover. This also occurs in more conventional oil and gas fields that are being “infilled” (where new wells are drilled in spaces between existing wells), and in “tight sands” fields (where fluid minerals are so distributed that dense well spacing is more profitable).89 Additionally, coal-bed methane is acquired by pumping water away from subsurface coal seams, thereby reducing pressure on coal-trapped methane and allowing its collection. This results in huge amounts of surplus water—in 2003 alone, more than half a billion barrels of water were coproduced by Wyoming coal-bed methane wells, mostly in the Powder River basin. That water must be disposed of in some manner. To understate the obvious, disposal of excess water is a problem with a rather short history in the west. Several options are available, including simple surface runoff, pressurized aerial spraying (forced evaporation), and well reinjection, but each entails potentially harmful effects to soil, vegetation, streams, and wildlife. The potential for harm is increased when coproduced water is high in inorganic salts, which initial information suggests will be true for water pumped off of some Red Desert coal seams.90 Coal-bed methane is already being extracted in several Red

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/21/08

10:25 AM

Page 163

Desert locations, such as Atlantic Rim southwest of Rawlins and the Black Buttes region southeast of Rock Springs.91 Most of these fields are slated for expansion, and many, many new fields are being planned. Conclusion Debates over the management of the Red Desert, as over wild land anywhere, commonly center on the shared goal of integrating commercial development with conservation of wildlife. However, claiming such a goal is deceptively simple, potentially disingenuous, and often used by resource developers, managers, researchers, and conservationists alike to dodge the fundamental truth: more commercial extraction of natural resources from the Red Desert, conducted in the current fashion, will continue to harm wildlife. Although the specific effects of individual developments on local wildlife populations can always be researched ad nauseam, the ultimate effect will be more distress, disruption, and displacement of more wildlife. If resource development in the Red Desert continues at its current pace and with its current methods, the “integration” so often alluded to will largely involve documenting the continued retreat of native species. The history of the Red Desert makes it clear that wildlife can retreat in different ways. Extirpation in the style of bison, grizzly bear, black-footed ferrets, and big-river fish like the humpback chub is the most obvious. But more subtle losses occur too. The wild character of remaining species can be stealthily eroded, just as some elk herds that previously migrated to the Red Desert are now artificially fed on designated feed grounds where animals are routinely captured, tested for disease, and slaughtered if infected92—approximating livestock more and free-ranging wildlife less. Elk still occupy the Red Desert, but their mass migrations are gone; pronghorn are common, but they can no longer drift across the Interstate 80 corridor; gray wolves have recently returned, but they will remain only occasional and furtive visitors into the foreseeable future.93 American society is aware of the potential for

wildlife extinctions and has reacted with such constructs as the Endangered Species Act. But it is less aware of the steady removal of the wild from its wildlife, and the consequences of that. Far-sighted individuals in the late nineteenth and early twentieth centuries reversed a staggering decline in some Red Desert species with actions that were unpopular then but today are hailed as wise and proper. Some commercial forces now operating in the Red Desert, such as the coal industry, are willing, with encouragement, to apply responsible and effective techniques of mitigation, reclamation, and environmental stewardship. And there are many people working hard to move the fundamental question into the center of the debate: is a wild Red Desert, with a complete complement of wild life, worth more to society than short-term economic gains? Generations are judged by their moral, aesthetic, and economic legacies, and over time the first two gain in importance over the third. This is how we now judge past caretakers of the Red Desert—we vilify the last hide hunters but celebrate the first conservationists—and there is no reason to expect that we will be judged any differently.

Vertebrate Wildlife of the Red Desert

163

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 164

Notes 1. “Fauna” refers collectively to all animals occupying a certain landscape. 2. All climatic estimates and comparisons are derived from interpolated U.S.-wide maps of climate data from the Daymet U.S. Data Center, http://www.daymet.org. 3. Endothermic animals, as birds and mammals, are sometimes casually termed “warm-blooded” in reference to their physiological ability to maintain near-constant internal temperatures. This contrasts with ectothermic (also poikilothermic), or “cold-blooded,” animals such as fish, amphibians, reptiles, and insects, whose body temperatures fluctuate with environmental temperatures. 4. J. B. Findley and S. Anderson, “Zoogeography of the montane mammals of Colorado,” Journal of Mammalogy 37 (1956): 575–577. See also G. P. Beauvais, “Mammal responses to forest fragmentation in the Central and Southern Rocky Mountains,” in Forest Fragmentation in the Southern Rocky Mountains, edited by R. L. Knight, F. W. Smith, S. W. Buskirk, W. H. Romme, and W. L. Baker (University of Colorado Press, Boulder, 2000), 179 –201. 5. Documented occurrences on file at the Wyoming Natural Diversity Database, University of Wyoming, Laramie. 6. G. T. Baxter and M. Stone, Amphibians and Reptiles of Wyoming, 2nd ed. (Wyoming Game and Fish Department, Cheyenne, 1985). R. C. Stebbins, A Field Guide to Western Amphibians and Reptiles, 3rd ed. (Houghton Mifflin, Boston, 2003). 7. Taxonomy and local distribution follows C. A. Long, The Mammals of Wyoming (University of Kansas Museum of Natural History Publications, Lawrence, 1965); as updated by E. R. Hall, The Mammals of North America, 2nd ed. (John Wiley and Sons, New York, 1981); and T. W. Clark and M. Stromberg, Mammals in Wyoming (University of Kansas Museum of Natural History Publications, Lawrence, 1987). Continental ranges estimated from B. D. Patterson, G. Ceballos, W. Sechrest, M. F. Tognelli, T. Brooks, L. Luna, P. Ortega, I. Salazar, and B. E. Young, Digital Distribution Maps of the Mammals of the Western Hemisphere, version 1.0 (NatureServe, Arlington, Va., 2003). 8. Clark and Stromberg, Mammals in Wyoming, 128, 138. For a general discussion of wildlife adaptations to cold deserts, see R. Fautin, “Arid country animals,” Wyoming Wildlife 42, no. 2 (1978): 21–23. 9. G. A. Hammerson, Amphibians and Reptiles in Colorado, 2nd ed. (University Press of Colorado, Niwot; Colorado Division of Wildlife, Denver, 1999), 73–74 and 82 – 85. Baxter and Stone, Amphibians and Rep164

tiles, 22 –27. 10. D. Knight, Mountains and Plains: The Ecology of Wyoming Landscapes (Yale University Press, New Haven, Conn., 1994), 87– 88. 11. As used here, “hibernate” refers broadly to true deep hibernation as well as seasonal metabolic torpor. 12. A. O. Cerovski, “Birds of the Red Desert,” this volume. 13. “Riparian” technically refers to being of or next to a water body. “Riparian vegetation” generally refers to the lush growth of plants that occurs near surface water. 14. Cerovski, “Birds.” 15. Knight, Mountains and Plains, 99. 16. G. P. Jones, “Sagebrush,” this volume. 17. An ungulate is a hoofed mammal. The term is commonly used, as here, to refer to big game such as pronghorn, mule deer, and elk. 18. J. H. Powell, “Distribution, habitat use patterns, and elk response to human disturbance in the Jack Morrow Hills, Wyoming” (master’s thesis, University of Wyoming, 2003). 19. Beauvais, “Mammal responses.” 20. R. D. Dorn, The Wyoming Landscape 1805 –1878 (Mountain West Publishing, Cheyenne, 1986), 86 – 88. 21. G. P. Beauvais and R. Smith, “A model of breeding habitat of the mountain plover (Charadrius montanus) in western Wyoming,” Western North American Naturalist 63 (2003): 88 –96. 22. For example, see Clark and Stromberg, Mammals in Wyoming, 153. 23. D. C. Pavlacky Jr., “Avian community ecology in juniper woodlands of southwestern Wyoming: Patterns of landscape and habitat utilization” (master’s thesis, University of Wyoming, 2003). 24. J. A. Lockwood, “Insects of the Red Desert: An exercise in scientific humility,” this volume. 25. J. W. Connelly, K. P. Reese, and M. A. Schroeder, “Monitoring of greater sage-grouse habitats and populations,” University of Idaho College of Natural Resources Experiment Station Bulletin 80, contribution no. 979 (Moscow, 2003). M. S. Drut, W. H. Pyle, and J. A. Crawford, “Technical note: Diets and food selection of sage grouse chicks in Oregon,” Journal of Range Management 47 (1994): 90 –93. G. D. Johnson and M. S. Boyce, “Feeding trials with insects in the diet of sage grouse chicks,” Journal of Wildlife Management 54 (1990): 89 –91. 26. For a complete and very readable discussion of North American fauna during the past 65 million years, see T. Flannery, The Eternal Frontier: An Ecological History of North America and Its People (Atlantic Monthly Press, New York, 2001). See also D. N. Walker, “Late Pleistocene/Holocene environmental changes in Wyoming: The mammalian

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 165

record,” in Late Quaternary Mammalian Biogeography and Environments of the Great Plains, edited by R. W. Graham, H. A. Semken Jr., and M. A. Graham (Illinois State Museum Scientific Papers 22, 1987), 334–392. The Jurassic period occurred 206 –144 million years before present, the Eocene 55–34 million years before present, and the Pleistocene 1.8 million to 10,000 years before present. Pleistocene climates are well described by B. Mears Jr., “Periglacial wedges and the late Pleistocene environment of Wyoming’s intermontane basins,” Quaternary Research 15 (1981): 171–198; C. Whitlock, “Postglacial vegetation and climate of Grand Teton and southern Yellowstone national parks,” Ecological Monographs 53 (1993): 173–198; and D. Gardner, “Environmental change in the Wyoming Basin’s Red Desert,” this volume. 27. The standard notation “bp” refers to “before present.” 28. B. Kurten and E. Anderson, Pleistocene Mammals of North America (Columbia University Press, New York, 1980), 319. E. C. Pielou, After the Ice Age: The Return of Life to Glaciated North America (University of Chicago Press, Chicago, 1991), 256. D. Gardner, “Horses come to the Red Desert,” this volume. 29. S. Mithen, After the Ice (Harvard University Press, Cambridge, Mass., 2004), 290 –291. B. Fagan, The Little Ice Age: How Climate Made History 1300–1850 (Basic Books, New York, 2000). S. T. Gray, S. T. Jackson, and J. L. Betancourt, “Tree-ring based reconstructions of interannual to decadal scale precipitation variability for northeastern Utah since 1226 A.D.,” Journal of the American Water Resources Association 40 (2004): 947–960. Knight, Mountains and Plains, 17–18. Gardner, “Environmental change.” 30. C. Parmesan and G. Yohe, “A globally coherent fingerprint of climate change impacts across natural systems,” Nature 421 (2003): 37– 42. T. L. Root, J. T. Price, K. R. Hall, S. H. Schneider, C. Rosenzweig, and J. A. Pounds, “Fingerprints of global warming on wild animals and plants,” Nature 421 (2003): 57– 60. 31. Walker, “Environmental changes.” 32. Gardner, “Horses.” 33. D. Flores, “Bison ecology and bison diplomacy: The southern plains from 1800 –1850,” Journal of American History 78 (1991): 465– 485. 34. S. Boyle and S. Owens, North American Beaver (Castor canadensis): A Technical Conservation Assessment (USDA Forest Service Rocky Mountain Region, Golden, Colo., 2007), 22 –25. See also R. Olson and W. A. Hubert, Beaver: Water Resources and Riparian Habitat Manager (University of Wyoming, Laramie, 1994). 35. Beaver were nearly extirpated from the entire state of Wyoming by about 1860; from ibid.

36. Knight, Mountains and Plains, 56 –59. Gardner, “Environmental change.” 37. Gardner, “Environmental change.” 38. USDI Fish and Wildlife Service, “Endangered and threatened wildlife and plants: 12-month finding for a petition to list the Columbian sharp-tailed grouse as threatened,” U.S. Federal Register 65 (2000): 60391– 60396. O. Oedekoven, “Columbian sharp-tailed grouse population distribution and habitat use in south-central Wyoming” (master’s thesis, University of Wyoming, Laramie, 1985). 39. In 1869 there were 90,000 cattle and sheep in Wyoming alone; this rose to 2,013,000 in 1889, and peaked at 6,788,000 in 1909. From C. M. Greenquist, “The American pronghorn antelope in Wyoming: A history of human influences and management” (Ph.D. diss., University of Oregon, Eugene, 1983). 40. “It seemed bad that the country should have sheep. They destroyed the grass, flowers, and little grouse in their nests on the ground with their sharp little hoofs.” From R. Wheeler, God’s Mountain: The Last of the Hahn’s Peak Homesteaders (Maverick Publications, Bend, Oreg., 2001), 110. 41. J. McPhee, Rising from the Plains (Farrar Straus Giroux, New York, 1986), 7–9. Knight, Mountains and Plains, 115. 42. N. Blair, The History of Wildlife Management in Wyoming (Wyoming Game and Fish Department, Cheyenne, 1987), 1–28. D. F. Lott, American Bison: A Natural History (University of California Press, Berkeley, 2002), 170 –179. 43. R. W. Koucky, “The buffalo disaster of 1882,” North Dakota History: Journal of the Northern Plains 50 (1983): 23–30. 44. M. Sandoz, The Buffalo Hunters: The Story of the Hide Men (University of Nebraska Press, Lincoln, 1978), 46 – 47. 45. Flores, “Bison ecology”; Koucky, “Buffalo disaster.” 46. R. F. Hosness and N. M. Frost, A Wyoming Bighorn Sheep Study, Wyoming Game and Fish Department Bulletin 1 (Cheyenne, 1942). 47. M. B. Boyce, R. K. Clark, and D. J. Jessop, “Recent advances in the diagnosis and treatment of psoroptic scabies in bighorn sheep,” Biennial Symposium of North American Wild Sheep and Goat Council 7 (1990): 125– 128. H. K. Buechner, “The bighorn sheep in the United States: Its past, present, and future,” Wildlife Monographs 4 (1960). 48. Greenquist, “American pronghorn antelope.” 49. H. P. Danz, Of Bison and Man: From the Annals of a Bison Yesterday to a Refreshing Outcome from Human Involvement with America’s Most Valiant of Beasts (University of Colorado Press, Boulder, 1997), 144. 50. D. A. Dary, The Buffalo Book: The Full Saga of the American Animal

Vertebrate Wildlife of the Red Desert

165

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 166

(Swallow Press, Athens, Ohio, 1989), 261. Dary also cites W. T. Hornaday’s declaration of the extinction of all wild bison in the United States (presumably excluding Yellowstone National Park) in the mid-1880s, and his warning that false rumors of wild remnants will persist. 51. T. A. Larson, History of Wyoming, 2nd ed. (University of Nebraska Press, Lincoln, 1965), 336 –369. Blair, History of Wildlife Management, 51, 61. 52. Knight, Mountains and Plains, 104, 118. 53. G. T. Baxter and M. D. Stone, Fishes of Wyoming (Wyoming Game and Fish Department, Cheyenne, 1995), 266 –271. Blair, History of Wildlife Management, 30. 54. Blair, History of Wildlife Management, 29 –55. 55. Ibid. 56. Ibid., 35. 57. Ibid., 54. 58. United Press International, “Poachers eliminate Red Desert buffalo,” Wyoming Tribune-Eagle, February 10, 1971. 59. W. J. Allred, “Re-establishment of seasonal elk migration,” Transactions of the North American Wildlife Conference 15 (1950): 597– 611. C. M. Cromley, “Historical elk migrations around Jackson Hole, Wyoming,” Yale School of Forestry and Environmental Studies Bulletin 104 (2000): 53– 65. J. Berger, “The last mile: How to sustain long-distance migration in mammals,” Conservation Biology 18 (2004): 320 –331. 60. H. Sawyer, R. Nielson, D. Strickland, and L. McDonald, 2005 Annual Report—Sublette Mule Deer Study Phase II: Long-Term Monitoring Plan to Assess Potential Impacts of Energy Development on Mule Deer in the Pinedale Anticline Project Area (Western Ecosystems Technology, Cheyenne, 2005). H. Sawyer and F. Lindzey, Jackson Hole Pronghorn Study (University of Wyoming Cooperative Fish and Wildlife Research Unit, Laramie, 2000). H. Sawyer, F. Lindzey, and D. McWhirter, “Mule deer and pronghorn migration in western Wyoming,” Wildlife Society Bulletin 33 (2005): 1266 –1273. 61. Berger, “The last mile.” 62. Blair, History of Wildlife Management, 31. Cromley, “Historical elk migrations.” 63. USDI Fish and Wildlife Service and USDI National Park Service, Draft Bison and Elk Management Plan and Environmental Impact Statement for the National Elk Refuge/Grand Teton National Park/John D. Rockefeller, Jr., Memorial Parkway (2005), http://www.fws.gov/bisonandelk plan/, 1–10. Cromley, “Historical elk migrations.” Blair, History of Wildlife Management, 44– 45 and 136 –138. 64. Cromley, “Historical elk migrations.” 166

65. More recent citizen-led campaigns have sought to designate portions of the Red Desert as a “wild horse refuge,” a “North American antelope range,” and various forms of statutory wilderness. The most recent proposal for designation as a national conservation area is described at http://www.reddesert.org/events/nca.php. See also Berger, “The last mile”; and C. Beers, “The Jack Morrow Hills,” Wyoming Wildlife 66, no. 7 (2002): 30 –39. 66. D. Henderson, “The sands,” Wyoming Wildlife 19, no. 7 (1955): 4– 6. D. Lockman, “Mama sage’s elk herd,” Wyoming Wildlife 42, no. 2 (1978): 25–27. H. Sawyer and R. Nielson, Seasonal Distribution and Habitat Use Patterns of Elk in the Jack Morrow Hills Planning Area, Wyoming (Western Ecosystems Technology, Cheyenne, 2005). 67. Knight, Mountains and Plains, 104–105. 68. Baxter and Stone, Fishes. “Lacustrine” refers to things related to lakes, as differentiated to those related to flowing rivers and streams. 69. M. K. Young, R. N. Schmal, T. W. Kohley, and V. G. Leonard, Conservation Status of Colorado River Cutthroat Trout, USDA Forest Service General Technical Report RM-GTR-282 (Fort Collins, Colo., 1996). M. R. Bower, “Distributions and habitat associations of bluehead suckers, flannelmouth suckers, and roundtail chubs in the upper Muddy Creek Watershed of southern Carbon County, Wyoming” (master’s thesis, University of Wyoming, 2005). 70. Blair, History of Wildlife Management, 175–176. 71. Baxter and Stone, Fishes, 270. 72. Blair, History of Wildlife Management, 176 –177. 73. J. F. Kamler and W. B. Ballard, “A review of native and nonnative red foxes in North America,” Wildlife Society Bulletin 30 (2002): 370 –379. 74. Members of this lobby include, for example, the American Wild Horse Preservation Campaign, the American Mustang and Burro Association, and Horse Rescue of America. 75. USDI Fish and Wildlife Service, Black-Footed Ferret Recovery Plan Status Report (Black-Footed Ferret Conservation Center, Fort Collins, Colo., 2002). 76. Dorn, Wyoming Landscape, 87– 88. 77. Hosness and Frost, Wyoming Bighorn Sheep Study. Boyce et al., “Recent advances.” 78. S. C. Trombulak and C. A. Frissell, “Review of ecological effects of roads on terrestrial and aquatic communities,” Conservation Biology 14 (2000): 18 –30. For a listing and review of the major ecological effects, both direct and indirect, of roads on wildlife in the Red Desert region, see C. Weller, J. Thomson, P. Morton, and G. Aplet, Fragmenting Our Lands: The Ecological Footprint from Oil and Gas Development; A Spatial Analysis

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 167

of a Wyoming Gas Field (Wilderness Society, Washington D.C., 2002); and usdi Bureau of Land Management, Draft EIS for the Pinedale Anticline Oil and Gas Exploration and Development Project, Sublette County, Wyoming (usdi Bureau of Land Management Pinedale Field Office, Pinedale, Wyo., 1999). 79. F. Ingelfinger, “The effects of natural gas development on sagebrush steppe passerines in Sublette County, Wyoming” (master’s thesis, University of Wyoming, Laramie, 2001). 80. A. G. Lyon, “The potential effects of natural gas development on sage grouse (Centrocercus urophasianus) near Pinedale, Wyoming” (master’s thesis, University of Wyoming, Laramie, 2003). 81. Powell, “Elk response to human disturbance.” Sawyer and Nielson, Seasonal Distribution. Also H. Sawyer, R. Nielson, F. Lindzey, L. Keith, J. Powell, and A. Abraham, “Habitat selection of Rocky Mountain elk in a nonforested environment,” Journal of Wildlife Management 71 (2007): 868 – 874. 82. Sawyer et al., 2005 Annual Report. 83. See discussion and citations in Beauvais, “Mammal responses.” 84. For example, see H. Sawyer and B. Rudd, Pronghorn Roadway Crossings: A Review of Available Information and Potential Options (Western Ecosystems Technology, Cheyenne, 2005). 85. D. Crowe and D. Strickland, “Weathering winter,” Wyoming Wildlife 48, no. 6 (1984): 20 –27. 86. W. Gasson, “Update: Red Rim,” Wyoming Wildlife 48, no. 3 (1984): 26 –30. Blair, History of Wildlife Management, 269 –272. 87. Blair, History of Wildlife Management, 128. 88. Weller et al., Fragmenting Our Lands. 89. For photos and discussion of a tight-sands gas field in the upper Green River valley, see ibid. 90. University of Wyoming Ruckelshaus Institute of Environment and Natural Resources, Water Production from Coalbed Methane Development in Wyoming: A Summary of Quantity, Quality, and Management Options (University of Wyoming, Laramie, 2005). 91. Ibid. 92. Wyoming Brucellosis Coordination Team, Report from the Wyoming Brucellosis Coordination Team to Governor Dave Freudenthal (2005). Wyoming Game and Fish Department, Elk Test-and-Removal Pilot Project Year Two Report (Cheyenne, 2007). 93. Wyoming Game and Fish Department, Final Wyoming Gray Wolf Management Plan (Cheyenne, 2003). Wyoming Game and Fish Department, Draft Wyoming Gray Wolf Management Plan (Cheyenne, 2007).

Bibliography Allred, W. J. “Re-establishment of seasonal elk migration.” Transactions of the North American Wildlife Conference 15 (1950): 597– 611. Baxter, G. T., and M. D. Stone. Amphibians and Reptiles of Wyoming. 2nd ed. Wyoming Game and Fish Department, Cheyenne, 1985. ———. Fishes of Wyoming. Wyoming Game and Fish Department, Cheyenne, 1995. Beauvais, G. P. “Mammal responses to forest fragmentation in the Central and Southern Rocky Mountains.” In Forest Fragmentation in the Southern Rocky Mountains, edited by R. L. Knight, F. W. Smith, S. W. Buskirk, W. H. Romme, and W. L. Baker. University of Colorado Press, Boulder, 2000. Beauvais, G. P., and R. Smith. “A model of breeding habitat of the mountain plover (Charadrius montanus) in western Wyoming.” Western North American Naturalist 63 (2003): 88 –96. Beers, C. “The Jack Morrow Hills.” Wyoming Wildlife 66, no. 7 (2002): 30 –39. Berger, J. “The last mile: How to sustain long-distance migration in mammals.” Conservation Biology 18 (2004): 320 –331. Blair, N. The History of Wildlife Management in Wyoming. Wyoming Game and Fish Department, Cheyenne, 1987. Bower, M. R. “Distributions and habitat associations of bluehead suckers, flannelmouth suckers, and roundtail chubs in the upper Muddy Creek Watershed of southern Carbon County, Wyoming.” Master’s thesis, University of Wyoming, 2005. Boyce, M. B., R. K. Clark, and D. J. Jessop. “Recent advances in the diagnosis and treatment of psoroptic scabies in bighorn sheep.” Biennial Symposium of North American Wild Sheep and Goat Council 7 (1990): 125–128. Boyle, S., and S. Owens. North American Beaver (Castor canadensis): A Technical Conservation Assessment. usda Forest Service Rocky Mountain Region, Golden, Colo., 2007. Buechner, H. K. “The bighorn sheep in the United States: Its past, present, and future.” Wildlife Monographs 4 (1960). Cerovski, A. O. “Birds of the Red Desert.” In this volume. Clark, T. W., and M. Stromberg. Mammals in Wyoming. University of Kansas Museum of Natural History Publications, Lawrence, 1987. Connelly, J. W., K. P. Reese, and M. A. Schroeder. “Monitoring of greater sage-grouse habitats and populations.” University of Idaho College of Natural Resources Experiment Station Bulletin 80, contribution

Vertebrate Wildlife of the Red Desert

167

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 168

no. 979. Moscow, 2003. Cromley, C. M. “Historical elk migrations around Jackson Hole, Wyoming.” Yale School of Forestry and Environmental Studies Bulletin 104 (2000): 53– 65. Crowe, D., and D. Strickland. “Weathering winter.” Wyoming Wildlife 48, no. 6 (1984): 20 –27. Danz, H. P. Of Bison and Man: From the Annals of a Bison Yesterday to a Refreshing Outcome from Human Involvement with America’s Most Valiant of Beasts. University of Colorado Press, Boulder, 1997. Dary, D. A. The Buffalo Book: The Full Saga of the American Animal. Swallow Press, Athens, Ohio, 1989. Daymet U.S. Data Center. http://www.daymet.org/. Dorn, R. D. The Wyoming Landscape 1805 –1878. Mountain West Publishing, Cheyenne, 1986. Drut, M. S., W. H. Pyle, and J. A. Crawford. “Technical note: Diets and food selection of sage grouse chicks in Oregon.” Journal of Range Management 47 (1994): 90 –93. Fagan, B. The Little Ice Age: How Climate Made History 1300–1850. Basic Books, New York, 2000. Fautin, R. “Arid country animals.” Wyoming Wildlife 42, no. 2 (1978): 21– 23. Findley, J. B., and S. Anderson. “Zoogeography of the montane mammals of Colorado.” Journal of Mammalogy 37 (1956): 575–577. Flannery, T. The Eternal Frontier: An Ecological History of North America and Its People, Atlantic Monthly Press, New York, 2001. Flores, D. “Bison ecology and bison diplomacy: The southern plains from 1800 –1850.” Journal of American History 78 (1991): 465– 485. Gardner, D. “Environmental change in the Wyoming Basin’s Red Desert.” In this volume. ———. “Horses come to the Red Desert.” In this volume. Gasson, W. “Update: Red Rim.” Wyoming Wildlife 48, no. 3 (1984): 26 – 30. Gray, S. T., S. T. Jackson, and J. L. Betancourt. “Tree-ring based reconstructions of interannual to decadal scale precipitation variability for northeastern Utah since 1226 A.D.” Journal of the American Water Resources Association 40 (2004): 947–960. Greenquist, C. M. “The American pronghorn antelope in Wyoming: A history of human influences and management.” Ph.D. diss., University of Oregon, Eugene, 1983. Hall, E. R. The Mammals of North America. 2nd ed. John Wiley and Sons, New York, 1981.

168

Hammerson, G. A. Amphibians and Reptiles in Colorado. 2nd ed. University Press of Colorado, Niqor: Colorado Division of Wildlife, Denver, 1999. Henderson, D. “The sands.” Wyoming Wildlife 19, no. 7 (1955): 4– 6. Hosness, R. F., and N. M. Frost. A Wyoming Bighorn Sheep Study. Wyoming Game and Fish Department Bulletin 1, Cheyenne, 1942. Ingelfinger, F. “The effects of natural gas development on sagebrush steppe passerines in Sublette County, Wyoming.” Master’s thesis, University of Wyoming, Laramie, 2001. Johnson, G. D., and M. S. Boyce. “Feeding trials with insects in the diet of sage grouse chicks.” Journal of Wildlife Management 54 (1990): 89 –91. Jones, G. P. “Sagebrush.” In this volume. Kamler, J. F., and W. B. Ballard. “A review of native and nonnative red foxes in North America.” Wildlife Society Bulletin 30 (2002): 370 – 379. Knight, D. Mountains and Plains: The Ecology of Wyoming Landscapes. Yale University Press, New Haven, Conn., 1994. Koucky, R. W. “The buffalo disaster of 1882.” North Dakota History: Journal of the Northern Plains 50 (1983): 23–30. Kurten, B., and E. Anderson. Pleistocene Mammals of North America. Columbia University Press, New York, 1980. Larson, T. A. History of Wyoming. 2nd ed. University of Nebraska Press, Lincoln, 1965. Lockman, D. “Mama sage’s elk herd.” Wyoming Wildlife 42, no. 2 (1978): 25–27. Lockwood, J. A. “Insects of the Red Desert: An exercise in scientific humility.” In this volume. Long, C. A. The Mammals of Wyoming. University of Kansas Museum of Natural History Publications, Lawrence, 1965. Lott, D. F. American Bison: A Natural History. University of California Press, Berkeley, 2002. Lyon, A. G. “The potential effects of natural gas development on sage grouse (Centrocercus urophasianus) near Pinedale, Wyoming.” Master’s thesis, University of Wyoming, Laramie, 2003. McPhee, J. Rising from the Plains. Farrar Straus Giroux, New York, 1986. Mears, B., Jr. “Periglacial wedges and the late Pleistocene environment of Wyoming’s intermontane basins.” Quaternary Research 15 (1981): 171–198. Mithen, S. After the Ice. Harvard University Press, Cambridge, Mass., 2004.

Gary P. Beauvais

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 169

Oedekoven, O. “Columbian sharp-tailed grouse population distribution and habitat use in south-central Wyoming.” Master’s thesis, University of Wyoming, Laramie, 1985. Olson, R., and W. A. Hubert. Beaver: Water Resources and Riparian Habitat Manager. University of Wyoming, Laramie, 1994. Parmesan, C., and G. Yohe. “A globally coherent fingerprint of climate change impacts across natural systems.” Nature 421 (2003): 37– 42. Patterson, B. D., G. Ceballos, W. Sechrest, M. F. Tognelli, T. Brooks, L. Luna, P. Ortega, I. Salazar, and B. E. Young. Digital Distribution Maps of the Mammals of the Western Hemisphere. Version 1.0. NatureServe, Arlington, Va., 2003. Pavlacky, D. C., Jr. “Avian community ecology in juniper woodlands of southwestern Wyoming: Patterns of landscape and habitat utilization.” Master’s thesis, University of Wyoming, 2003. Pielou, E. C. After the Ice Age: The Return of Life to Glaciated North America. University of Chicago Press, Chicago, 1991. Powell, J. H. “Distribution, habitat use patterns, and elk response to human disturbance in the Jack Morrow Hills, Wyoming.” Master’s thesis, University of Wyoming, 2003. Root, T. L., J. T. Price, K. R. Hall, S. H. Schneider, C. Rosenzweig, and J. A. Pounds. “Fingerprints of global warming on wild animals and plants.” Nature 421 (2003): 57– 60. Sandoz, M. The Buffalo Hunters: The Story of the Hide Men. University of Nebraska Press, Lincoln, 1978. Sawyer, H., and F. Lindzey. Jackson Hole Pronghorn Study. University of Wyoming Cooperative Fish and Wildlife Research Unit, Laramie, 2000. Sawyer, H., F. Lindzey, and D. McWhirter. “Mule deer and pronghorn migration in western Wyoming.” Wildlife Society Bulletin 33 (2005): 1266 –1273. Sawyer, H., and R. Nielson. Seasonal Distribution and Habitat Use Patterns of Elk in the Jack Morrow Hills Planning Area, Wyoming. Western Ecosystems Technology, Cheyenne, 2005. Sawyer, H., R. Nielson, F. Lindzey, L. Keith, J. Powell, and A. Abraham. “Habitat selection of Rocky Mountain elk in a nonforested environment.” Journal of Wildlife Management 71 (2007): 868 – 874. Sawyer, H., R. Nielson, D. Strickland, and L. McDonald. 2005 Annual Report — Sublette Mule Deer Study Phase II: Long-Term Monitoring Plan to Assess Potential Impacts of Energy Development on Mule Deer in the Pinedale Anticline Project Area. Western Ecosystems Tech-

nology, Cheyenne, 2005. Sawyer, H., and B. Rudd. Pronghorn Roadway Crossings: A Review of Available Information and Potential Options. Western Ecosystems Technology, Cheyenne, 2005. Stebbins, R. C. A Field Guide to Western Amphibians and Reptiles. 3rd ed. Houghton Mifflin, Boston, 2003. Trombulak, S. C., and C. A. Frissell. “Review of ecological effects of roads on terrestrial and aquatic communities.” Conservation Biology 14 (2000): 18 –30. United Press International. “Poachers eliminate Red Desert buffalo.” Wyoming Tribune-Eagle, February 10, 1971. University of Wyoming Ruckelshaus Institute of Environment and Natural Resources. Water Production from Coalbed Methane Development in Wyoming: A Summary of Quantity, Quality and Management Options. University of Wyoming, Laramie, 2005. usdi Bureau of Land Management. Draft EIS for the Pinedale Anticline Oil and Gas Exploration and Development Project, Sublette County, Wyoming. usdi Bureau of Land Management Pinedale Field Office, Pinedale, Wyo., 1999. usdi Fish and Wildlife Service. “Endangered and threatened wildlife and plants: 12-month finding for a petition to list the Columbian sharp-tailed grouse as threatened.” U.S. Federal Register 65 (2000): 60391– 60396. ———. Black-Footed Ferret Recovery Plan Status Report. Black-Footed Ferret Conservation Center, Fort Collins, Colo., 2002. usdi Fish and Wildlife Service and usdi National Park Service. Draft Bison and Elk Management Plan and Environmental Impact Statement for the National Elk Refuge/Grand Teton National Park/John D. Rockefeller, Jr., Memorial Parkway. 2005. http://www.fws.gov/ bisonandelkplan/. Walker D. N. “Late Pleistocene/Holocene environmental changes in Wyoming: The mammalian record.” In Late Quaternary Mammalian Biogeography and Environments of the Great Plains, edited by R. W. Graham, H. A. Semken Jr., and M. A. Graham, Illinois State Museum Scientific Papers 22, 334–392. 1987. Weller, C., J. Thomson, P. Morton, and G. Aplet. Fragmenting Our Lands: The Ecological Footprint from Oil and Gas Development; A Spatial Analysis of a Wyoming Gas Field. Wilderness Society, Washington, D.C., 2002. Wheeler, R. God’s Mountain: The Last of the Hahn’s Peak Homesteaders. Maverick Publications, Bend, Oreg., 2001.

Vertebrate Wildlife of the Red Desert

169

Proulx_04_pp.121-170:Layout 1

8/7/08

12:52 PM

Page 170

Whitlock, C. “Postglacial vegetation and climate of Grand Teton and southern Yellowstone national parks.” Ecological Monographs 53 (1993): 173–198. Wyoming Brucellosis Coordination Team. Report from the Wyoming Brucellosis Coordination Team to Governor Dave Freudenthal. 2005. Wyoming Game and Fish Department. Draft Wyoming Gray Wolf Management Plan. Cheyenne, 2007. ———. Elk Test-and-Removal Pilot Project Year Two Report. Cheyenne, 2007. ———. Final Wyoming Gray Wolf Management Plan. Cheyenne, 2003. Wyoming Natural Diversity Database. University of Wyoming, Laramie. Young, M. K., R. N. Schmal, T. W. Kohley, and V. G. Leonard. Conservation Status of Colorado River Cutthroat Trout. usda Forest Service General Technical Report RM-GTR-282. Fort Collins, Colo., 1996.

170

Gary P. Beauvais

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 171

6. Birds of the Red Desert Andrea Orabona

o the casual observer, the vast, rolling expanse of sagebrush punctuated by isolated sky islands of rock that forms the long view of the Red Desert may appear as an endless wasteland, a suggested moonscape, a biological desert. My son verbalized this view on his first trip to the Red Desert when I asked for his impression: “It looks like a barren wasteland that stretches as far as I can see.” Spend some time exploring this unique landscape, however, and you will gain an entirely different perspective. My vocation as a wildlife biologist and my avocation as a wildlife watcher give me ideal opportunities to observe birds in the Red Desert. Every summer, I conduct a Breeding Bird Survey (bbs) in the heart of the greater Red Desert. A bbs route consists of fifty stops, one-half mile apart, at which all birds seen and heard during a three-minute time frame are recorded.1 Mine is one of 108 routes in Wyoming, and one of thousands conducted by volunteers each year throughout the United States and Canada. The bbs is the longest-running survey of its kind and tracks population changes of hundreds of species of birds whose behavior and habitat preferences during the breeding season (April through July for most birds) lend themselves well to detection during the survey. When given the chance, I also enjoy personal forays into the

T

Red Desert to find solace in the open space, unique beauty, and rich birdlife that the desert has to offer. With each season’s change comes a shift in both the number and mix of bird species that inhabit the Red Desert. There are about fifty species of birds that live within the sea of sagebrush that covers much of the desert’s surface.2 Some—the Greater Sage-Grouse, the Sage Thrasher, the Sage Sparrow, and Brewer’s Sparrow—are called sagebrush obligates because they rely on sagebrush for nearly all of their nesting, feeding, and shelter needs (see Table 6.1 for the scientific names of species mentioned in this chapter). Many others, such as Vesper and Lark sparrows, Horned Larks, Green-tailed Towhees, Common Poorwills, and the occasional Loggerhead Shrike, dwell amid the mixed desert shrubs and grasses interspersed with the sagebrush. There are a handful—Golden Eagles, Prairie Falcons, Ferruginous Hawks, Common Ravens, Cliff Swallows, Rock Wrens, and Say’s Phoebes—that key in on the cliff walls, rims, and ledges of the sky islands, buttes, and rocky outcrops. The desert’s streams, lakes, and ponds, always a welcome sight in this arid landscape, hold an unimaginable number of ducks, including Mallards, Northern Pintails, and all three teals—Green-winged, Blue-winged, and Cinnamon. The abundance of birds tucked within the springs and seeps, stands of aspen, and pockets of willows scattered across the desertscape is 171

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 172

more than one would first imagine—American Robins, Warbling Vireos, Yellow-rumped Warblers, House Wrens, Western WoodPeewees, Western Tanagers, American Kestrels, Great Horned Owls, and White-breasted Nuthatches, just to name a few. Some are expected but still a pleasure to find, and others are a treat when you discover them. Seeing each of these species is a treasure that makes me feel very alive and, like the changes we see in plants throughout the passage of each year, marks the shifting of the seasons in the desert. Spring Spring is the season of abundant moisture in the Red Desert as winter’s snow and ice slowly yield to mild daytime temperatures and spring rains dampen the land. It is also a season rich in the diversity of birds found across this varied landscape. With warming days, copious water, blossoming flowers, and an awakening of insects come the return of the migrants—those species that breed in the United States and Canada during our summer season and depart at the end of the breeding cycle for warmer climates during winter. Some arrive in the Red Desert in the spring and stay to nest and raise their young, while others rely on the desert as a stopover site to rest and refuel before continuing their journey to breeding grounds farther north. Red-winged Blackbirds, particularly the males, push the spring season a bit. I have been astonished by a few hardy souls seen flying over the Red Desert in early March when the ground is still completely blanketed with snow. As soon as the marshy meadows and pond edges thaw, however, their hoarse song — konk-la-reeee—rings out as one of the true harbingers of spring. Other members of the blackbird family—Brewer’s Blackbird, the Common Grackle, and the Western Meadowlark—arrive a little later and stake their claims to the moist meadows and open, grassy shrublands. Bullock’s Orioles, also in the blackbird family, return to patches of mature aspen that ring the natural springs and seeps around places like Steamboat Mountain and Bush Rim. 172

In these groves, they suspend their pendulous nests from twigs and small branches. One of Wyoming’s claims to fame is that it hosts more than half of the entire Greater Sage-Grouse population of about 140,000 birds that occur in parts of eleven western states and Alberta, Canada.3 Although Greater Sage-Grouse live in Wyoming year-round, they have specific seasonal ranges within the Red Desert (and elsewhere). Like the turning of calendar pages, each segment of the population predictably returns in early spring to its own historical lek, or display ground. There the males fan their tails, swish their wings, puff out their breast feathers, and repeatedly inflate and deflate the two air sacs on their chests to create a garbled, popping sound. During the crack-of-dawn hours in March and April, the males perform this unusual strutting ritual on more than two hundred leks throughout the Red Desert, vying for a chance to mate with as many females as possible during spring courtship.4 Even if you are not an early riser, observing this amazing display is worth sacrificing some sleep at least once in your lifetime. There is only a brief window of opportunity to view a few species in the Red Desert during spring migration. The American Redstart, a small wood-warbler, stops to regain energy in the desert’s isolated stands of aspen and dogwood. Although it pauses only a day or two before moving on to breeding grounds farther north, the male’s striking black-and-orange plumage will forever remain in your memory. You may spot an occasional Great Blue Heron patiently waiting for a chance to nab a small fish or frog from one of the creeks temporarily swollen with spring runoff, or a small flock of American White Pelicans as it wheels around to land on a desert lake to rest. The Broad-tailed Hummingbird can startle you with a dynamic appearance as it follows the bloom of wildflowers northward from its wintering grounds in Latin America. Because of its wings’ need for speed, you’ll typically hear the buzzing of a hummingbird before seeing it as it searches out tubular, nectar-producing wildflowers and protein-rich insects. Although some broad-tails may stay and nest within the

Andrea Orabona

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 173

Table 6.1. Bird species mentioned in this chapter Common name

Scientific name

Common name

Scientific name

American Avocet American Coot American Crow American Goldfinch American Kestrel American Redstart American Robin American White Pelican American Wigeon Black-billed Magpie Black-capped Chickadee Blue-gray Gnatcatcher Blue-winged Teal Brewer’s Blackbird Brewer’s Sparrow Broad-tailed Hummingbird Bullock’s Oriole Burrowing Owl California Gull Canada Goose Canvasback Cinnamon Teal Clark’s Nutcracker Cliff Swallow Common Grackle Common Poorwill Common Raven Dark-eyed Junco Ferruginous Hawk Gadwall Golden Eagle Great Blue Heron Greater Sage-Grouse Great Horned Owl Green-tailed Towhee Green-winged Teal Horned Lark

Recurvirostra americana Fulica americana Corvus brachyrhynchos Carduelis tristis Falco sparverius Setophaga ruticilla Turdus migratorius Pelecanus erythrorhynchos Anas americana Pica hudsonia Poecile atricapillus Polioptila caerulea Anas discors Euphagus cyanocephalus Spizella breweri Selasphorus platycercus Icterus bullockii Athene cunicularia Larus californicus Branta canadensis Aythya valisineria Anas cyanoptera Nucifraga columbiana Petrochelidon pyrrhonota Quiscalus quiscula Phalaenoptilus nuttallii Corvus corax Junco hyemalis Buteo regalis Anas strepera Aquila chrysaetos Ardea herodias Centrocercus urophasianus Bubo virginianus Pipilo chlorurus Anas crecca Eremophila alpestris

House Wren Killdeer Lark Sparrow Lazuli Bunting Least Flycatcher Least Sandpiper Lesser Yellowlegs Loggerhead Shrike Long-billed Curlew Long-eared Owl MacGillivray’s Warbler Mallard Marbled Godwit Mountain Bluebird Mountain Plover Northern Flicker Northern Harrier Northern Pintail Northern Shoveler Orange-crowned Warbler Pinyon Jay Prairie Falcon Red-breasted Nuthatch Red-tailed Hawk Red-winged Blackbird Ring-billed Gull Rock Wren Rough-legged Hawk Sage Thrasher Sage Sparrow Savannah Sparrow Say’s Phoebe Semipalmated Sandpiper Short-eared Owl Snowy Owl Spotted Sandpiper Steller’s Jay

Troglodytes aedon Charadrius vociferus Chondestes grammacus Passerina amoena Empidonax minimus Calidris minutilla Tringa flavipes Lanius ludovicianus Numenius americanus Asio otus Oporornis tolmiei Anas platyrhynchos Limosa fedoa Sialia currucoides Charadrius montanus Colaptes auratus Circus cyaneus Anas acuta Anas clypeata Vermivora celata Gymnorhinus cyanocephalus Falco mexicanus Sitta canadensis Buteo jamaicensis Agelaius phoeniceus Larus delawarensis Salpinctes obsoletus Buteo lagopus Oreoscoptes montanus Amphispiza belli Passerculus sandwichensis Sayornis saya Calidris pusilla Asio flammeus Bubo scandiacus Actitis macularius Cyanocitta stelleri

Birds of the Red Desert

173

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 174

Table 6.1 continued Common name

Scientific name

Tree Swallow Vesper Sparrow Violet-green Swallow Warbling Vireo Western Meadowlark Western Tanager Western Wood-Peewee White-breasted Nuthatch White-faced Ibis Willet Wilson’s Phalarope Wilson’s Warbler Yellow-rumped Warbler Yellow Warbler

Tachycineta bicolor Pooecetes gramineus Tachycineta thalassina Vireo gilvus Sturnella neglecta Piranga ludoviciana Contopus sordidulus Sitta carolinensis Plegadis chihi Catoptrophorus semipalmata Phalaropus tricolor Wilsonia pusilla Dendroica coronata Dendroica petechia

aspen stands close to water, their nests are tiny and difficult to find. However fleeting they may be, these moments of rest and replenishment are critical for migrating birds. If the roads are passable, a visit to the Killpecker Dunes in April or May is a must. There you will find dunal ponds—seasonal ponds in the sand dunes — that harbor migrating shorebirds and waterbirds. The exact location of these ponds each spring is random, for they are created by slowly melting snow and ice that is trapped under the sand dunes during winter, and snowfall in the desert is never easy to predict. American Avocets, Marbled Godwits, Willets, Spotted Sandpipers, and Wilson’s Phalaropes find these dunal ponds and gorge themselves on insects, invertebrates, and small crustaceans found in the damp mud or on the water’s surface. With a good spotting scope or a steady hand on your binoculars, you can discern the different bill sizes and shapes and the corresponding feeding tactics of the assorted shorebirds. Watch as the American Avocet sweeps its upturned bill back and forth to stir up insects and crustaceans from 174

the bottom of the pond. The long bill of the Marbled Godwit and shorter bill of the Willet both probe mud but do so at different depths, so they exploit the aquatic invertebrates found throughout the layers of mud without directly competing with one another for food. As it races back and forth near the water’s edge, the Spotted Sandpiper needs only its small bill to pick up insects from the mud’s surface. Wilson’s Phalaropes use a completely different feeding strategy and are comical to watch as they turn in tight circles on the water’s surface, stirring up a banquet from the bottom of the pond. As the days lengthen and the north-facing slopes of Steamboat Mountain begrudgingly give up their last patches of snow, these seasonal ponds also disappear. With the ponds go the shorebirds, completing their spring migration to more stable aquatic areas elsewhere. Although several species of shorebirds nest in Wyoming, most are migrants and only a few are known to nest in the Red Desert. Yet without a stop at the temporary dunal ponds to rest and refuel, the completion of this annual sojourn for these species would be much more difficult.

Andrea Orabona

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 175

Summer If you dodge the seasonal rainstorms and visit the Red Desert in June, you will find the wild iris in bloom and the birds transformed into their signature breeding colors and full of song. Birds that inhabit the desert during the breeding season usually return to the same territories to nest—and often the same tree, shrub, cliff, or rock outcrop— especially if they successfully raised young the previous year. This provides a great opportunity to detect your quarry, because birds are very predictable during the breeding season. Anyone with an interest can identify different species of birds by their discrete songs and calls, their specific coloration, and their behavior. On record, there are at least a dozen species of raptors, or birds of prey, that nest in the Red Desert.5 A few, like the Golden Eagle and Great Horned Owl, remain year-round. Most are migrants, though, and their return coincides with an increase in activity of their preferred foods—mainly rodents, plus a few birds, snakes, lizards, and insects to round out the diet. Owls arrive in February and March to find mates and raise their young. Although owl pairs will often reuse the same nest each year, they prefer to use the abandoned or alternate nests of other species rather than build their own. (Short-eared and Snowy owls are the exception, and they build nests on the ground.) Great Horned and Long-eared Owls find suitable stick nests in the secluded stands of pines atop Steamboat and Green mountains or in the scattered patches of aspen around the desert’s seeps and springs. The Burrowing Owl, one of my favorites because of its unique behavior, is adapted to spending part of its life underground. Mates choose the shelter and safety of abandoned prairie dog or ground squirrel burrows for nesting and raising their young. If you find a Burrowing Owl pair or small nesting colony and approach their burrows too closely, you may be surprised to hear the sound of a rattlesnake instead of a raptor. The Burrowing Owl’s ability to imitate the sound of an agitated rattler is a tactic that keeps potential predators at bay. In mid- to late summer,

when adults and young are aboveground and active during the day, set up your spotting scope in the Killpecker Dunes or near Brannan Reservoir and enjoy the antics of a Burrowing Owl clan as they bob their heads, fling dirt with their feet, and flutter around their burrows. Most of the eagles, hawks, and falcons are migrants that arrive in March and April to begin the annual ritual of courtship, nest building, and raising young. Some of the most noticeable nests are made by Golden Eagles, one of the Red Desert’s largest avian inhabitants and one of a handful of raptors that stays in Wyoming year-round. Look for their impressive stick nests on cliff ledges around Oregon Buttes, Steamboat Mountain, north Table Rock, Green Mountain, the Jack Morrow Hills, and the Boar’s Tusk. Once the young have grown and left the nest, hike up to a flat spot below it and check out the remains of what young Golden Eagles are raised on. The striking Ferruginous Hawk, with its reddish back, wings, legs, and wing linings and its snowy white underparts, also builds a conspicuous stick nest on a rocky outcrop or on the ground at the edge of a hillside. (It builds stick nests in large trees too, but these are scarce in the desert, while hillsides and rock outcrops are plentiful.) Like many raptors, Ferruginous Hawks often return to the same nest every spring. Each year, the pair add more sticks to their nest until, over time, it resembles a miniature volcano — the larger the volcano, the older the nest. Wyoming records for the Ferruginous Hawk include more than a thousand nests throughout the entire greater Red Desert.6 No matter where you may find them, keep in mind that Ferruginous Hawks are especially sensitive to human disturbance during the nesting season and will abandon their nest, eggs, or young if a disturbance crosses a boundary that may be invisible to us. Binoculars or a spotting scope can bring your view right into a nest if you have a good vantage point. Red-tailed Hawks are one of the most abundant raptors in Wyoming. Like many of the hawks, their body color can vary from a light phase to a dark phase, making identification a challenge, but

Birds of the Red Desert

175

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 176

their red tails are diagnostic. Redtails build sturdy stick nests wherever trees grow—in the clumps of tall aspen that crop up near water and in the pines around Green and Steamboat mountains— but they can be found hunting for rodents, birds, reptiles, and insects in the surrounding open country. Although their nests may be well hidden among tree branches, look for the revealing tail color—light rust to vibrant crimson—as they soar in the thermals. Formerly called the Marsh Hawk, the Northern Harrier is an easy-to-identify raptor that nests on the ground around the desert’s marshes, moist meadows, and seeps. The gray male and brown female both sport a bright white rump patch and usually hover or soar low over the ground in search of their next meal. Young harriers look completely different from adults, with their stunning burnt orange plumage. Regardless of their color, the adults and young alike have rather owlish-looking faces because of the specialized ruff of feathers around them. For owls and harriers both, this facial disk acts like a giant reflector, directing the sound of prey to the birds’ ears during the hunt. Like Golden Eagles, Prairie Falcons choose cliff ledges on which to nest. Unlike Golden Eagles, however, Prairie Falcons build no nest to speak of, laying their eggs in a small depression directly on the rocky ledge instead. Prairie Falcon eggs are very pointed at one end, so if disturbed, they roll in a circle rather than off the cliff edge—a very effective adaptation. Although nests can be hard to find, look for the sandy brown flash of a Prairie Falcon as it hunts for small birds, mammals, insects, and lizards around the Honeycomb Buttes, along Bush Rim, and near Continental Peak. The vivid colors, bold patterns, and hover-and-dive hunting behavior of the American Kestrel can be found in the desert where Northern Flickers have chiseled nesting cavities in mature pine and aspen trees. Although the kestrel is Wyoming’s smallest falcon, these birds announce the presence of their nest when approached too closely with a loud, sharp klee-klee-kleeklee-klee-klee-klee, making them sound anything but small. Waterfowl— ducks and geese—return to the Red Desert in March and April when the aquatic plants and insects they eat 176

flourish. At first glance, you would never imagine that so many marshes and bodies of water dot the desert landscape. Most ducks and geese use these oases during spring and autumn migration, while others—Mallards, Gadwalls, Green-winged Teals, Northern Pintails, and Northern Shovelers—stay to nest and raise their broods in these special places. April also brings the return of gulls, waterbirds (like pelicans), and marsh birds (herons and their ilk), followed by shorebirds in May. Although most move on and nest elsewhere, each species is an important component of the desert, adding richness and diversity to the landscape. The handful that do stay to nest and raise their young—like the Killdeer, the Mountain Plover, and the Spotted Sandpiper— choose different niches within the desert to do so. For example, Killdeer can adapt to nearly any habitat type as long as there are graveled areas near water for nesting. Mountain Plovers, on the other hand, prefer dry, open country with lowgrowing plants and patches of bare ground, and, unlike Killdeer, they never nest near water. Although not as closely related to Killdeer, Spotted Sandpipers are typically found near water, but they choose to nest along the shoreline. This strategy is not accidental; rather, it allows each species to exploit available nesting and foraging sites while curtailing competition for limited resources. The Greater Sage-Grouse, a year-round inhabitant of the Red Desert, has become the avian equivalent of a poster child for keeping intact the vast, healthy, unbroken expanses of the sagebrush habitat it requires for all phases of its life. As a sagebrush obligate species, sage-grouse rely on sagebrush plants for at least 90 percent of their food, nesting, and shelter requirements, as well as the associated spring, seep, and wet meadow microhabitats interspersed throughout the Red Desert. These moist environments provide an important source of water and produce copious amounts of insects sage-grouse chicks need during their first formative weeks of life. Some biologists believe that, by providing all the habitat components required by an “umbrella” species like the Greater Sage-Grouse, the habitat needs of many other species

Andrea Orabona

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 177

will also be provided. For instance, the sagebrush-obligate songbirds — Sage Thrasher, Sage Sparrow, and Brewer’s Sparrow — return from northern and central Mexico to nest each summer in sagebrush habitat in the Red Desert and across much of Wyoming. Like the Greater Sage-Grouse, these species need large, unbroken expanses of sagebrush, and all the pieces of the habitat puzzle that go along with it, to successfully nest and raise their young. Hence, if land stewards ensure that there is enough habitat to sustain Greater Sage-Grouse, these other sagebrushobligate species will also be taken care of under the same management umbrella. Like the Greater Sage-Grouse, a few other species—Mountain Plovers, Ferruginous Hawks, and Burrowing Owls—are also receiving a growing amount of national and international attention. Fortunately, these species are still holding their own throughout most of the Red Desert because human disturbance, for the most part, has been limited. All of these species are sensitive in some way to human-caused changes and disturbance. Because most individual birds of these species return to the same sites to nest every year, and Wyoming hosts most or much of the habitat these species require to thrive, we have an obligation to protect the breeding and foraging habitats of all these species within the Red Desert. Although some species, like Red- and White-breasted nuthatches, Black-capped Chickadees, and Dark-eyed Juncos, eke out a living year-round in the desert’s protected pockets of trees and shrubs, most of the songbirds that breed in the Red Desert arrive in spring when the insects, flower buds, nectar, and seeds on which they feed abound. As varied as their names, this large and assorted group includes blackbirds, flycatchers, gnatcatchers, swallows, sparrows, buntings, warblers, vireos, thrushes, towhees, and wrens. With this kind of avian smorgasbord, the summer season is a wonderful time to revel in the grand medley of color and sound that songbirds bring to the desert during this intense time of year. The lives of some songbirds could be a traveler’s dream. Orange-crowned Warblers, for instance, spend our winter months in

Guatemala, while Yellow Warblers winter in the Bahamas, northern Mexico, Peru, Bolivia, and the Amazon.7 Each year, they return to the Red Desert to nest and raise their young in the scattered patches of aspen and willows, for they are two of the many riparian-obligate species. Simply put, these are species that rely on the layers of trees, shrubs, grasses, and wildflowers that grow around water for nearly all of their food, water, shelter, and nesting needs. Many others, such as House Wrens, Warbling Vireos, Western Wood-Peewees, American Goldfinches, MacGillivray’s Warblers, Lazuli Buntings, and Tree Swallows, are called ripariandependent species. Again, they are found during the summer months in the trees and shrubs of the Red Desert’s riparian areas, relying on this habitat to satisfy many, but not all, of their needs. Whether a riparian obligate or dependent, all of these species migrate to the Neotropics, or New World tropics, of Mexico, Central America, South America, and the Caribbean for the nonbreeding season and return each year to these oases in the Red Desert. Whether you travel by vehicle or foot, a stop at any of the riparian areas in the Red Desert will yield a rewarding birding experience and a refreshing break from the summer’s heat. Find a shady spot with a good vantage point for bird-watching and you will be dazzled by the diversity of species in these small patches of habitat. At one unnamed spring in the Red Desert, I watched as Least Flycatchers, Western Wood-Peewees, and Western Tanagers swooped from their perches and caught insects in midair, a feeding behavior known as hawking. A turn of my head brought the vibrant blue of a male Lazuli Bunting into view, and in the opposite direction I spotted a pair of Yellow-rumped Warblers discreetly perched among the leaves. And all the while I could hear the exuberant songs of House Wrens and American Robins advertising their presence, as Vesper and Savannah sparrows sang from the moist, grassy edges of the spring. During any time of the day, check out the goings-on around the myriad rocky outcrops scattered across the desert’s landscape. Here you can enjoy the aerial acrobatics of Cliff and Violet-green swallows as they dive and maneuver after flying insects. Or pause

Birds of the Red Desert

177

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 178

and listen for the varied trills and buzzes of a male Rock Wren as he advertises his availability to females and defends his territory against rival males from a prominent rocky perch. With a keen ear, you can add a few more species to your list. During the summer months, Green-tailed Towhees sing almost nonstop during the day from their perches on junipers or tall shrubs. Listen for their characteristic song that seems to ask the listener to “drink up all your tea, now please.” Not nearly as common but also found in patches of juniper around places like north Table Rock are Blue-gray Gnatcatchers. Once heard, their thin, wheezy, buzzy song is diagnostic, even if you never get a good look as they dart and flit from one shrub to another in search of insects. There are many accessible places in the Red Desert during the summer, and each offers an element of surprise. One of my most intriguing Red Desert memories came at the end of a summer day when I took one final hike along Jack Morrow Creek. Around a corner and chiseled high in a tall bank along the creek were half a dozen cavities. I raised my binoculars, and there, peering back at me, was a Northern Flicker sitting on a nest in one of the cavities. I was used to seeing Mountain Bluebirds and Say’s Phoebes use these earthen spaces for nesting along the creeks and in the soft walls of the Honeycomb Buttes, but a woodpecker was an unexpected sight. This was a great example of an individual’s ability to adapt to local conditions in its ultimate quest to pass on its genes. Autumn In the Red Desert and elsewhere in the United States and Canada, most of the insect- and seed-eating songbirds, all marsh birds, most waterbirds, all shorebirds, and many of the raptors pack on the fat and travel south during autumn migration to avoid the stress of cold weather and to find a steady supply of food in the warmer climates of the southern United States, Mexico, Central and South America, and the Caribbean. Some are called long178

distance migrants because they travel thousands of miles and take weeks or months to arrive in their nonbreeding-season habitats. Others are short-distance migrants because they only need to go as far as the relatively warm coasts and inlands of California, Texas, Arizona, or northern Mexico to survive during our winter months. Different groups of birds leave at different times after the breeding season, affording more chances for bird-watchers to observe and enjoy them. Typically, most songbirds that travel longer distances to the tropics of Mexico and farther south will migrate from late August through September, while those that winter within the United States can wait until mid-September, October, and even into November to make the move. Other longdistance migrants must gather together and leave earlier too. For instance, shorebirds migrate from mid-August through September, and the long-distance hawks start their migratory flights in late August. Whenever their departure date, most migrants make the journey in flocks, and they rely on certain places in the Red Desert to gather together and fuel their bodies first. Some species may be a rare find in the desert during the summer months. Wilson’s Warblers, for example, prefer dense, higher-elevation thickets of willows and alders near water. After the breeding season, however, they move to lower elevations to gather together and fatten up on insects and berries before embarking on their migratory journey to Panama.8 Look for these small, yellow songbirds as they forage in the shrubs around Pickett and Scotty lakes. Other species, like the Mountain Plover, nest in the Red Desert but can be tough to see during the summer months, when they are dispersed across the landscape in their breeding territories. But come time for autumn migration in late August, and Mountain Plovers amass, or stage, on large, level expanses of land with low-growing plants, like Bush Rim and the flats around Bear Creek, before making their seasonal migratory flight to Baja and northern Mexico. Still other birds can be found in the Red Desert only during spring or autumn migration, which highlights the importance of the desert for many more species of birds than those we call

Andrea Orabona

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 179

breeders. While water levels in the Red Desert reservoirs are dictated by the amount of yearly rain and snowfall, Pickett, Scotty, and McKay are naturally occurring lakes within the Great Divide basin (where the continental divide splits and rejoins itself ) and are not as vulnerable to the fickle dictates of annual precipitation. These lakes vary in size, but all are extremely important and reliable feeding areas for shorebirds, marsh birds, waterbirds, and waterfowl during autumn migration. A languid afternoon spent behind a spotting scope at any of these natural desert lakes will yield a surprising number and variety of migrants as they stage, feed, and rest. On one memorable day spent looking for autumn migrants, I saw small flocks of Least, Western, Semipalmated, and Spotted sandpipers; a few Marbled Godwits, American Avocets, and Lesser Yellowlegs; some Long-billed Curlews, Whitefaced Ibises, and American White Pelicans; both California and Ring-billed gulls; and hundreds upon hundreds of waterfowl, including Canada Geese, American Coots, Canvasbacks, Northern Shovelers, American Wigeons, and Mallards. A thrilling sight, indeed, in the waning warmth of the season. Seasonal sightings like these illustrate the real connectivity between breeding grounds, wintering grounds, and the migratory corridors that join the two. For biologists, the innovative way to manage populations of birds extends beyond the borders of our own small pieces of this planet to cross state, regional, and even international boundaries. We in Wyoming share a precious resource with our neighbors in the birds that grace our landscapes, since they are “ours” for only part of each year. Autumn migration, when the birds of the Red Desert leave, and spring migration, when they return, accentuate this point best. Winter If snow conditions are just right to travel by snowmobile, crosscountry skis, or snowshoes, the stark beauty of winter in the Red Desert and the tenacity of its handful of year-round birds are humbling sights to behold.

As sagebrush obligates, Greater Sage-Grouse feed almost exclusively on the buds and leaves of sagebrush shrubs during the winter months. Dense stands of productive sagebrush that are tall enough to rise above the snow provide both winter food and cover for sage-grouse. Their cryptically colored feathers, mottled to blend in with their surroundings, and their habit of stalking through rather than flying over the shrubs can make sage-grouse difficult to spot. With some patience and a good vantage point from which to look into a stand of sagebrush, you can watch for movement within the stand as sage-grouse forage from shrub to shrub. Some of the more obvious species in the Red Desert are the year-round raptors, like the Golden Eagle and Great Horned Owl, and winter-season raptors, like the Rough-legged Hawk. Turn your eyes to the sky and look for these species as they search for jackrabbits, small mammals, and other sustenance in winter’s meager bounty. Common Ravens, American Crows, and Black-billed Magpies are what biologists call omnivorous species; that is, they eat a mixture of plants, animals, and insects. Some species, like the Blackcapped Chickadee, switch from insects to seeds when insects are no longer available. Both are ideal feeding strategies for winter inhabitants of the Red Desert because they enable birds to exploit the food sources that are available. Each of these species forms large winter flocks in the desert as they forage together by day and roost (or rest) together by night. This is a strength-in-numbers strategy that relies on more pairs of eyes to find food and more bodies to generate warmth during the harsh winter months. Although mature conifer trees (evergreens that produce cones) and junipers (evergreens that produce berries) are scarce in most places within the Red Desert, clusters of them can be found atop Green Mountain and Pine Canyon, as well as on some of the sky islands like Steamboat Mountain, Oregon Buttes, and north Table Rock. These isolated pockets of evergreens, along with stands of tall, dense sagebrush, provide shelter and important year-round food resources for those species able to find enough

Birds of the Red Desert

179

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 180

sustenance to withstand the physiological stresses of winter. Foraging and roosting together in variable-sized flocks during the winter months are Steller’s Jays, with their deep blue body feathers and black crests; Pinyon Jays, which are similar in color but sport a shorter tail and lack the crest and dark head of the Steller’s; and Clark’s Nutcrackers with their gray, black, and white coloring. All three species feed on pine nuts and conifer seeds and even cache them during times of abundant nut and seed production to sustain themselves during leaner times. Nut and seed caches also enable stands of conifers and junipers to expand and are a bonanza for the small mammals that happen to find them. This is a truly symbiotic relationship—it benefits both the birds and the plants upon which they feed and sustains the existence of both. The variety of birds found in the Red Desert during the winter pales in comparison to that during the other seasons, but the ability of these creatures to survive is no small feat in an environment that can be harsh and unforgiving even during the best of times. From vast spaces to hidden places, we have come full circle through the seasons in the Red Desert, back to the awakening of the land in spring and the return of the rich tapestry of birdlife that abounds in this wondrous landscape. Professionally as a biologist and personally as a devotee of the natural world, I hope that you too will explore and marvel at the diversity of birds that this fascinating piece of our planet has to offer, for with understanding and appreciation comes a passion to protect these feathered treasures for future generations to enjoy.

180

Notes 1. North American Breeding Bird Survey, U.S. Geological and Canadian Wildlife Service, http://www.pwrc.usgs.gov/bbs/. 2. C. Paige and S. A. Ritter, Birds in a Sagebrush Sea: Managing Sagebrush Habitats for Bird Communities (Partners in Flight Western Working Group, Boise, Idaho, 1999). 3. National Wildlife Federation, http://www.nwf.org/wildlife/. 4. Sage Grouse Lek Locations throughout Wyoming, unpublished database (Wyoming Fish and Game Department, 2002). 5. A. O. Cerovski, B. Grenier, B. Oakleaf, L. Van Fleet, and S. Patla, Atlas of Birds, Mammals, Amphibians and Reptiles in Wyoming (Wyoming Game and Fish Department Nongame Program, Lander, 2004). 6. Green River Resource Area Raptor Database, unpublished database (Bureau of Land Management). 7. P. R. Erlich, D. S. Dobkin, and D. Wheye, The Birder’s Handbook (Simon and Schuster, New York, 1988). 8. Ibid.

Andrea Orabona

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 181

7. Horses Come to the Red Desert Dudley Gardner

he Adams-Onís Treaty of 1819 divided the Red Desert region between the United States and Spain. The dividing line ran along the 42nd parallel to the continental divide.1 What did Spain know of the Red Desert? Did Spanish influence extend as far north as the Wyoming Basin? Hints can be found in the diaries of explorers, Spanish maps, and the arrival of horses into one of the largest natural pastures in the interior west. When Coronado traveled north into New Mexico in 1540, he brought horses and iron trade goods into Pueblo villages with long-established trade networks extending north into present Colorado, Utah, and Wyoming, populated by Utes, Shoshonis, and Comanches. These three groups got horses through trade and raids in Pueblo and Spanish villages in northern New Mexico. The question here is, when did the Utes, Shoshonis, and Comanches acquire the horse?2 For fifty years following the arrival of Coronado, the Spanish had contact with the Pueblos along the Rio Grande. In 1598, when Juan de Oñate came to settle the area, he met Indians who could speak Spanish.3 From 1598 on, horses grazed on the grasses of the Rio Grande Valley. Soon Native Americans rode, raised, and traded these animals. They acquired some of these horses through trade. Others were stolen from the Spaniards or were lost stock.4 Natives from the north also raided settlements to get

T

mounts. As the Spanish at the missions tried to keep horses from the Indians, stealing became the best way to get them. Within a century after the horse reached the northern Rio Grande, the Utes and Comanches had horses. These animals would reach the Red Desert via an odyssey that began in New Mexico. The Spanish claim on southwest Wyoming from 1540 to 1821 and the history of the horse go hand in hand. The Southern Utes probably had indirect contact with the Spanish in the late sixteenth century and likely gained access to horses in the 1600s. The Ute Nation extended from just north of Taos, New Mexico, to the present southern border of Wyoming and west to the Great Basin in present Utah. Its northern border abutted the ShoshoniComanche homelands. It appears that the Shoshonis and closely related Comanches both lived at one time in present southwest Wyoming and obtained the horse from the Utes. The Comanches left the region for the southern plains.5 The Spanish reported Comanche presence in the southern plains in 1706.6 Their origins puzzled the Spaniards: “An ancient tradition, related to Brigadier Pedro de Rivera in 1726, placed their origin in a land called Teguayo, hundreds of miles northwest of Santa Fe.”7 There is disagreement over when the Comanches went south, and for good reason.8 If the Comanches, allied with the Utes, attacked Taos in 1706, it meant that the horse had reached the 181

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 182

Wyoming Basin in the late 1600s. This is not beyond the realm of possibility but defies some long-held conventions and requires substantiation. It is clear that between 1706 and 1776, when Domínguez and Escalante reached the Uintas, that the Yamparika Comanches who lived in the region had the horse. Tracking the spread of the horse into southwest Wyoming is difficult. At one time the Eastern Shoshonis and the Yamparika Comanches both probably lived in the Wyoming Basin.9 Shoshoni and Comanche are both Numic languages. Ernest Wallace and E. Adamson Hoebel contend that Shoshoni and Comanche are mutually intelligible.10 Based on linguistic studies, these languages probably split sometime in the sixteenth or seventeenth century. Wallace and Hoebel point to the possibility of the Comanches originating in southwestern Wyoming or northwestern Colorado. Both share several common ceremonies, such as the Antelope Dance. The groups’ split may have occurred either shortly before or after the Comanches acquired the horse. As to when the horse reached the Wyoming Basin, John C. Ewers notes that Northern Shoshoni tradition “claims their kinsmen, the Comanche, furnished them their first horse.” Ewers goes on to contend that the Shoshonis probably acquired the horse shortly after 1705.11 If Ewers is correct and if the Shoshonis reached the northern plains before the arrival of La Vérendrye in the 1730s, it is possible that the horse reached southwestern Wyoming in the early 1700s. In the seventeenth century, millennium-old trade networks were conduits for metal goods and horses to move from New Mexican settlements north into the interior. On the northern plains existed a trade system paralleling the trade networks that fanned out from Pueblo villages along the Rio Grande. On the plains, trade networks radiated from farming villages along the Missouri River and its tributaries. For centuries Native people had served as middlemen and -women in a trade system that linked rivers, plains, mountains, and woodlands. When the French reached the northern plains in the 1700s, they dealt with these traders. Both the Pawnees (in Nebraska) and the Mandans (in present North Dakota) did business in the western interior. The Pawnee and 182

Mandan villages had access to French goods. Kettles and other iron implements from these villages reached the Taos Indians in the late 1600s.12 By 1700 the southwest traders at Pecos and Taos were actively engaged in trading European goods with the Pawnees and the Mandans over long distances. Juan de Oñate brought breeding mares north from Santa Fe, and by 1598, horses were on the upper Rio Grande.13 Governor Don Diego de Vargas, of Nuevo Mexico, notes on June 4, 1691, “My duty is conquest and reduction, rather than chasing off and destroying a rancheria and a few bands of Indians who try to steal horses by the light of the moon.”14 Stolen horses went into the interior of North America in exchange for slaves and hides. Horses reached the north by decades before the Spaniards. Early in the eighteenth century, Ute horse raids in the valley of Taos along the northern Rio Grande alarmed Spanish officials. The Indians at Taos Pueblo complained of being harassed by the Utes. In 1719, Spain declared war on the Utes. In August of that year, ten Ute raiders ran off four horses and kidnapped one boy. The children Utes kidnapped could be used as servants or traded for horses. Their raids resulted in a rapid increase in the size of their horse herds. One accusation held that they committed “robberies of horse herds every day.”15 Alejo Martin, a settler at Rio Arriba in northern New Mexico in 1719, testified “that he is of the feeling armed forces should be directed against the Indians of the Ute nation since, although up to the present, they have shown themselves friends of the Spaniards, yet under this friendship they had carried off some animals.”16 In fact, through raids on Pueblo villages and trade, the Utes’ horse herd had grown so large that they now traded these horses to neighboring tribes. Among those they traded with were the Comanches and their cousins the Shoshonis. The Comanches also raided horse herds, and this tribe, with relatives in the Wyoming Basin, was seen as a direct threat to security of the Spanish frontier.17 Governor Antonio de Valverde’s 1719 campaign against the Utes and Comanches provided excellent details regarding the land north and east of Taos. On September 21, Valverde reported see-

Dudley Gardner

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 183

ing “many stone landmarks placed by . . . the Apache tribe so that they may not lose themselves.”18 Valverde’s expedition is significant in two ways: Valverde saw horses well out on the plains along the Arkansas River; and he firmly plants the Comanches in contact with the Apaches along that river. In 1719 he notes that the Pawnees had standing trading relations with the French and, as a result, had obtained guns. In fact, the Pawnees had shot and wounded an Apache whom Valverde encountered on the Arkansas River in present southern Colorado.19 The Apaches, according to Valverde’s report, had settled along the Arkansas to get away from the Pawnees and the French.20 By 1719 the Apaches had been pushed south by both the Pawnees and the Comanches. The Comanches had horses; the Pawnees, guns.21 In 1720 Colonel Don Pedro de Villasur extended Spanish knowledge of the region north to the Platte River.22 The colonel neared southeastern Wyoming by traveling up the Platte into present Nebraska.23 Villasur went north looking for Frenchmen and Pawnees, wanting to confirm whether or not the French had indeed settled among the Pawnees.24 Part of the Spanish concern grew out of the appearance of French metal axes and guns, trade items that reached New Mexico in 1719.25 That French goods had moved so far into the interior alarmed the Spanish government in Santa Fe. The Spanish rulers there concluded that the Pawnees, and possibly the Utes and Comanches, had allied with the French.26 Mounted on horses, the Utes and Comanches roamed free throughout the mountains and plains. The Pawnees, with relatively few horses, were at a disadvantage. This would change. In at least one case in 1720 the Pawnees bought two horses with two French guns.27 Soon, they too had become mounted threats to Spanish interests. The Pawnees, Comanches, and Utes all had horses by 1720, and horses had extended north to the boundaries of present Colorado and Wyoming. Spanish documents indicate that the Pawnees, Comanches. and Apaches who lived along the South and North Platte had horses.28 In the late 1730s the Mandans told

the La Vérendrye brothers that tribes to the south of them traveled by horse.29 With both the Utes and the Comanches hunting and traveling the mountains and plains of present Colorado by 1720, and with both still having relatives in the Wyoming Basin, there is the possibility the horse had reached into the interior mountain basins by the early eighteenth century.30 Supporting this is that European metal and ceramic artifacts dating to the eighteenth century have been found in archaeological excavations in southwest Wyoming. The presence of European or Mexican manufactured goods dating to the 1700s adds to a growing body of data that indicate an eighteenth-century arrival of the horse in the Wyoming Basin.31 By 1738, tribes of the northern plains had knowledge of Spaniards and horses. In 1738 the Mandans told explorer Pierre Gaultier de Varennes, Sieur de La Vérendrye, that the Arikaras (Panana) and Pananis (Pawnee) “had some horses like the whites.” Louis-Joseph and François, two sons of La Vérendrye, explored the northern plains in 1743. An Assiniboin, in 1738 or 1739, had informed the elder La Vérendrye that the previous summer he had “killed one of the people, who was covered with iron.” He added, “If I had not first killed the horse, I would not have got the man.”32 The Gens des Chevaux (Horse People) appeared as mounted travelers on the plains by the time of La Vérendrye’s travels from Quebec. And mounted Gens de Serpent (People of the Serpent, or Shoshonis) had begun raiding on the plains on horseback. La Vérendrye provides a description for the Gens de Serpent in 1741. He notes they are “considered very brave.” It is said, he continues, “that in 1741 they completely defeated seventeen villages, killing all the men and the old women, making slaves of the young women, and selling them on the seacoast for horses and certain merchandise.”33 By the time Americans reached the interior west, horses grazed on the plains and in the mountains. The horse, by 1800, was part of the fauna of the American west. Lewis and Clark, in their travels through the northern plains and Rockies, kept an eye out for horse herds. Their stories of trad-

Horses Come to the Red Desert

183

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 184

ing for horses gives us a solid indication of how far north horses had extended in the intermountain west in the early 1800s.34 In July of 1805, Lewis and Clark saw horses at the three forks of the Missouri,35 where the Madison, Gallatin, and Jefferson rivers come together. Almost due south, across the Yellowstone Plateau and at the tip of the Wind River range, is the Red Desert. The horses that had originated in the American southwest had reached the northern plains and Rockies well before Lewis and Clark arrived in 1804. It can surely be assumed that tribes in present Wyoming had horses in the 1700s. The size of the northern plains’ herds in 1804 adds to this assumption. At Camp Fortunate in present southwest Montana, Lewis purchased three horses. This point is west and north of the Wyoming Basin. It is clear that when Lewis traded for these animals in 1805, horses were living deep into the American interior. Lewis writes on August 18, 1805, that he obtained horses from the “Shoshoni.” He gave “a uniform coat, a pair of leggings, a few handkerchiefs, three knives, and some other small articles” for three very good horses. Lewis valued those trade items at twenty dollars “in the United States.”36 Clark adds: “Purchased of the Indians three horses for which we gave a Chiefs coat some handkerchiefs and shirt legins and a few arrow points and etc.”37 Horses from New Spain had bred and spread far beyond Spain’s control, north into present Canada.

has since been supported by such noted anthropologists as Raymond Wood and Margot Liberty.39 While it is difficult to piece together the first interactions between the Shoshonis and European trappers, it is interesting to note that half the rendezvous took place along the Green River or one of its tributaries.40 Hence, all of the trappers’ rendezvous “took place in lands claimed by the Shoshoni.”41 Although battles between trappers and Shoshonis took place, these appear to have been sporadic, and it is difficult to determine their cause. Generally, trappers’ speaking relations with the Shoshonis were friendly in nature. In addition to their favorable foreign policy toward Americans, the Shoshonis were traders, and trading required cordial relationships to ensure success.

Shoshoni Traders Some authors have suggested that the Shoshonis actually developed trading fairs prior to the early 1820s, when Europeans arrived. In his classic study of the horse in Blackfoot culture, John Ewers names southwestern Wyoming as one of the trading centers for the dissemination of the horse.38 It would make sense that the Shoshonis also traded other items at their trade fairs and that the development of the rendezvous was patterned on these Shoshoni fairs. This hypothesis was first put forth by Ewers and

184

Dudley Gardner

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 185

Notes 1. The Adams-Onís Treaty came about through negotiations between the then secretary of state John Quincy Adams and the Spanish foreign minister to the United States, Don Luis de Onís. The continental divide, unknown to most diplomats, split at the east end of the 42nd parallel. If the divide specified in the 1819 treaty meant the east end of the Great Divide basin, the border lay roughly north of present Rawlins. If the reference to the continental divide meant the west end of the basin, the international border lay northeast of present Rock Springs. The treaty was signed in Washington, D.C., on February 22, 1819. Spain ratified the agreement on October 24, 1820. The treaty technically entered into force on February 22, 1821, but changes in Mexico created some ambiguity. The Mexican Revolution in 1821 led to some uncertainty as to the terms of the treaty. The Adams-Onís Treaty was not ratified by the United States and the new Republic of Mexico until 1831. 2. According to Virginia McConnell Simmons, The Ute Indians of Utah, Colorado, and New Mexico (Boulder: University of Colorado Press, 2000), 29, the Utes took their first horses from Santa Fe in 1640. 3. Gasper Perez de Villagrá, A History of New Mexico (Albuquerque: Rio Grande Press, 1962), 74. This work, translated by Gilberto Espinosa, first appeared in Spain in 1610. Villagrá notes that Juan de Oñate, who founded Santa Fe, received permission from the crown to settle New Mexico in 1595. He took “possession” of the land in 1598 (ibid., 136). Near present Bernalillo, Villagrá writes, “three naked [men] approached; one of them, drawing near to the general, said in a loud voice in Spanish, ‘Thursday, Friday, Saturday, and Sunday.’ We were astounded to hear these words from their lips” (141, 142). These three “naked men” pointed them to a village where “Christian Indians whom Castaño de Sosa had brought with him from New Spain” lived. They had settled among the Pueblos and would serve as translators for Juan de Oñate (142). David J. Weber, The Spanish Frontier in North America (New Haven, Conn.: Yale University Press, 1992), 78 – 80, notes that in 1581 a party of three Franciscans, led by Fray Agustín Rodríguez and accompanied by seven soldiers, traveled to the Pueblos along the northern Rio Grande. In 1582 Antonio de Espejo went north to find Fray Rodríguez. The priest had been killed, and Espejo turned his energies toward finding silver and gold. He exaggerated his findings, and not until 1590 would a serious attempt at settlement occur. In 1590 Gaspar Castaño de Sosa, without obtaining permission from the king, led 170 men, women, and children up the Pecos and then into the Rio Grande Valley. Officials in Mexico sent Juan Morlete to arrest Castaño and return him and his followers to

Mexico City. Official settlement of New Mexico would come in 1898. 4. In A History of New Mexico, Villagrá notes that when he found a group of Spanish deserters, two of them had escaped the Indians, “abandoning the horses” (150). In another instance three men lost their horses, which they had to attempt to find. Again Villagrá does not indicate whether they found the horses (175). He does note that Indians helped with the horses in 1598. In one case they brought “great quantities of water for the horses” (167). In the battle for Acoma in 1598, the Acoma Indians “killed several . . . horses” (199). What all of this shows is that slowly horses were finding their way into Indian villages. 5. Stanley Noyes, Los Comanches: The Horse People, 1751–1845 (Albuquerque: University of New Mexico Press, 1993), xix. Noyes contends: “The Comanches were Shoshonis, who on obtaining horses, split off from their parent tribe in the seventeenth century, riding from the mountains of present Wyoming southeast into the Great Plains.” 6. Thomas W. Kavanagh, The Comanches: A History, 1706 –1875 (Lincoln: University of Nebraska Press, 1999), 1. 7. John L. Kessell, Spain in the Southwest: A Narrative History of Colonial New Mexico, Arizona, Texas, and California (Norman: University of Oklahoma Press, 2002), 225. See also Alfred Barnaby Thomas, trans., After Coronado: Spanish Exploration Northwest of New Mexico 1696 –1727, 2nd ed. (Norman: University of Oklahoma Press, 1966); and Colin G. Calloway, One Vast Winter Count: The Native American West before Lewis and Clark (Lincoln: University of Nebraska Press, 2003), 119 –211. 8. Noyes, Los Comanches, xix. 9. Fray Angelico Chavez, The Domínguez-Escalante Journal: Their Expedition through Colorado, Utah, Arizona, and New Mexico in 1776, 32 –33, 52. 10. Ernest Wallace and E. Adamson Hoebel, The Comanches: Lords of the South Plains (Norman: University of Oklahoma Press, 1952). 11. John C. Ewers, The Horse in Blackfoot Indian Culture (Washington, D.C.: Smithsonian Institution Press, 1980), 6. The question also needs to be asked: who is to say that the horses did not move north on their own? 12. See Thomas, After Coronado, 173, 246 –247. 13. Kessell, Spain in the Southwest, 75–76. 14. John L. Kessell and Rick Hendricks, eds., By Force of Arms: The Journals of Don Diego de Vargas, New Mexico, 1691–93 (Albuquerque: University of New Mexico Press, 1992), 59. 15. Thomas, After Coronado, 103, 105. Alfred Thomas translates a number of documents related to the Spanish expeditions out onto the western Great Plains in 1719 and 1720. His focus is on the expedition and

Horses Come to the Red Desert

185

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 186

the campaign against the Utes, Comanches, and Pawnees. He does, however, tell much about the inner workings of the Spanish government in New Mexico. In Marc Simmons, Spanish Government in New Mexico (Albuquerque: University of New Mexico Press, 1990), the author notes that in 1776 the problem of small bands of Indians stealing horses had grown so acute that the royal treasury “was forced to appropriate money for 1,500 horses to replace those stolen from the citizenry and soldiers” of northern New Mexico (8). 16. Thomas, After Coronado, 99 –102. 17. Ibid., 103, 105. Escalante is clear on this point. In 1776 he places the Yamparika Comanches in present northwest Colorado and parts of southwest Wyoming. Escalante contends that his Ute guide told him that the Green River separated the “Yuta [Utes] from the Comanche” (Chavez, Domínguez-Escalante Journal, 52). 18. Thomas, After Coronado, 111. 19. Ibid., 132. In one instance the Spanish in 1720 counted the horse herd of a camp of “sixty-nine Apache.” This camp had 680 animals (117). At another Apache camp the Spaniards counted 103 horses (118). These herds appear to have been along the Purgatory River, and in this same area the Spanish expedition reported that three Comanches and one Ute had stolen 20 horses and a colt from the Apaches (122). 20. Ibid., 102 –133. 21. Ibid., 251. 22. The Spanish called the North Platte the San Lorenzo (Thomas, After Coronado, 273). 23. Ibid., 135–137. Actually, paranoia gripped the government of Santa Fe, and, of course, it led to the presence of Frenchmen “two hundred leagues from Santa Fe” being exaggerated (148 –149). 24. Ibid., 219 –220. 25. Ibid., 246 –247. In this work the translator gives the “Declaration of Don Francisco Lorenzo de Cassados” to Governor Don Juan Domingo de Bustamante at Santa Fe, April 21, 1724. Francisco states he has seen “some French muskets that some of the native Indians carry, among those who come into this kingdom for peaceful barter” (247). While the records have some contradictions, it is clear that the Pawnees had guns and metal axes (cf. ibid., 173). 26. Kavanagh, in The Comanches, does not focus on the 1719 –1720 expedition but provides insights into the Comanches’ early history (63– 79). At one point he notes that the Wichita ethnonym for the Comanche was na’taa’, meaning “snake” (69). (An ethnonym is the name of a people or ethnic group.) He goes on to note that in 1739 two French brothers, Pierre and Paul Mallett, along with eight companions, traveled from 186

Illinois to New Mexico and encountered “friendly Comanches” they called the Lalitanes. Kavanagh suggests that the two terms “na’taa’“ and “Lalitanes” are the same (68 – 69). Intriguingly, the Shoshonis of southwestern Wyoming were also called Snakes. La Vérendrye called the Shoshonis “Gens du Serpent” (People of the Snake). He said that they were very brave and in 1741 completely defeated seventeen villages, “making slaves of the young women, and selling them for horses and certain merchandise.” W. Raymond Wood, ed., The Explorations of the La Vérendryes in the Northern Plains, 1738 – 43 (Lincoln: University of Nebraska Press, 1980), 107. 27. Thomas, After Coronado, 173. 28. Kavanagh, in The Comanche (63–79), places the Comanche in contact with Spanish settlements by 1706. 29. Wood, Explorations. 30. In 1765 Juan Maria Antonio Rivera led a small party from Santa Fe to the Dolores River in present southwest Colorado. Rivera noted that the Utes had horses in this area and gave some of the Native Americans he encountered horses for presents. This expedition confirms the spread of the horse into the northern interior by 1765. Joseph P. Sanchez, Explorers, Traders, and Slavers: Forging the Old Spanish Trail, 1678 –1850 (Salt Lake City: University of Utah Press, 1997), 13. 31. The number of eighteenth century historic artifacts found in the area is not large, but the diversity of items makes the prospect of Shoshoni and Ute access to European goods an intriguing possibility. 32. Wood, Explorations, 58, 64. Wood notes that there is some question of whether the Panama are Arikaras and the Pananis, Pawnees. It is apparent, however, that the Mandans were aware of horses and of the Spanish by 1738 and clearly by 1742. 33. Wood, Explorations, 107. See page 121 for discussion of the Gens de Serpents being Shoshonis, or Snake Indians. On page 99 there is a reference to the Gens de Serpents as being very brave. 34. Gary E. Moulton, ed., The Journals of the Lewis and Clark Expedition (Lincoln: University of Nebraska Press, 1988 –1991), vols. 3–7. 35. Ibid., vol. 4, 422 – 428, 430. 36. Ibid., vol. 5, 117. 37. Ibid., 118. 38. John C. Ewers, The Horse in Blackfoot Indian Culture (Washington, D.C.: Smithsonian Institution Press, 1980). 39. W. Raymond Wood and Margot Liberty, eds., Anthropology on the Great Plains (Lincoln: University of Nebraska Press, 1980), 98 –107. 40. Howard R. Lamar, The Trader on the American Frontier: Myth’s Victim (College Station: Texas A&M University Press, 1977). Lamar also

Dudley Gardner

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 187

comments on trade networks, stating: “From this bird’s-eye view of the vast native-white trading frontiers of the trans-Mississippi west, several basic facts should be clear: first, that exploitative Indian-Indian and Indian-white trade in the west did not begin with American fur traders and Santa Fe merchants, but preceded actual white contact by generations. Second, we should see that the horse and gun were not simply new tools of warfare and food gathering, but trade items themselves whose presence motivated warfare and intensified the trade impulse to the point that perhaps more human beings were sold into slavery or exploited for reason of trade than ever before. . . . Thus in our periodic re-examination of the frontier, we should push our accounts of trade back in time and show maps of prehistoric Indian trade centers and routes, and then depict the Spanish, the French, the British and the American ones” (25–26, 28). 41. Virginia Cole Trenholm and Maurine Carley, The Shoshonis: Sentinels of the Rockies (Norman: University of Oklahoma Press, 1964), 56.

Horses Come to the Red Desert

187

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 188

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 189

8. Insects of the Red Desert An Exercise in Scientific Humility Jeffrey A. Lockwood

n honest entomologist describing the insect fauna of the Red Desert has a great deal in common with Socrates. When the oracle at Delphi proclaimed that no man was wiser than Socrates, the philosopher set out to refute the gods by finding a wiser man than he. But after visiting with a venerated sage, Socrates recounted, “When I began to talk with him, I could not help thinking that he was not really wise, although he was thought wise by many, and wiser still by himself.” Disappointed in his encounter, Socrates went away muttering, “Well, although I do not suppose that either of us knows anything really beautiful and good, I am better off than he is—for he knows nothing, and thinks that he knows. I neither know nor think that I know.” This tale of Greek humility is a potent warning for those who might attempt to claim knowledge about the Red Desert’s insects — or perhaps any other group of living organisms in that vast landscape. To put the matter bluntly, insects are probably the most poorly known life-form; the Red Desert is the most poorly known of desert ecosystems in North America; and deserts are the most poorly known of the earth’s terrestrial biomes. We can even quantify, to some extent, the scope of our ignorance. Using hits on a Google search of the World Wide Web as a crude means of determining human interest, if not knowledge, we find that there are about 80 million sites referring to “ocean,”

A

56 million pertaining to “forest,” and just 27 million related to “desert.” Among the North American deserts, the greatest number of sites refer to the Mojave (1.6 million), followed by the Sonoran (1.4 million), the Great Basin (770,000), and the Chihuahuan (286,000). Wyoming’s Red Desert logs in with 240,000 hits— the least-known representative of the least-known ecosystem. As for organisms, the big winner is plants (51 million sites), followed by birds (32 million) and then by insects (14 million). This wouldn’t seem so bad in terms of our ignorance, except when we compare the amount of information on the Web to the biodiversity of various taxa. Mammals, for example, have 10 million sites (9 percent of all sites concerning the most common lifeforms) but account for less than 0.5 percent of biological diversity. Even reptiles and amphibians are overrepresented, being about as diverse as mammals but still having 2 – 4 percent of the websites. Insects, which represent about 70 percent of the life-forms on earth, account for just 12 percent of the hits. And, finally, using hits on a Google search of the Web for “insect” and “Red Desert” in March of 2006 as a crude means of determining human interest yielded just 314 hits,1 essentially none of which actually had information about the insects of the Red Desert other than the occasional site recommending ways of repelling insects when visiting this inhospitable place. In the hope 189

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 190

of finding more relevant information, a search for “insect fauna” and “Red Desert” produced a paltry 6 sites. Five of these were gigantic ecological reports, databases, or bibliographies in which the search words occurred somewhere on the site but not together. The only site that actually had something about the insect fauna of the Red Desert included a research brief about a butterfly that is found in lots of places—but not the Red Desert. A more sophisticated investigation of the available scientific literature using library databases turned up not a single paper or thesis dedicated to the insects of Wyoming’s Red Desert. This is not to say, of course, that no entomologist has ever ventured into this realm as part of a research project. However, it appears that nobody ever did so with the explicit purpose of assessing the biodiversity of the Red Desert in its own right. There are various reports of the insect fauna of shrublands of the western United States,2 although surveys are far less numerous than those of friendlier landscapes such as forests and meadows. The general sense of the scientific community seems to be that sagebrush and greasewood communities support a rather prosaic assemblage of insect species, although many of these creatures appear to be limited to such areas. By collecting across a broad range of desert shrub communities, Jorgenson and Tingey managed to amass some three hundred species of insects, which suggests a respectable but hardly impressive biological diversity. Such estimates accord with what one might expect from communities having very low levels of primary production (plant biomass) due to harsh climates and poor soils. However, at least one family of insects would, if it had a spokesman, argue passionately for a very different interpretation of the Red Desert. The richest biogeographic database for a group of insects in Wyoming was developed through a cooperative venture of the U.S. Department of Agriculture and the University of Wyoming. The Wyoming Grasshopper Information System was intended to track pest species and their population dynamics, but in the course of the statewide surveys from 1988 to 1992, any and all grasshopper species were collected.3 And a remarkable view of 190

the state’s insect diversity emerged. The 45,782 specimens represented 103 species of grasshoppers, none of which were exotic or introduced. There may be no larger family of animals in the western United States composed entirely of native species. To put the diversity of this single family of insects (in the United States there are 613 families, in the 28 orders that make up the class Insecta) into biological context, consider that the richness of grasshoppers rivals that of the entire class Mammalia, with 120 exotic and native species in Wyoming. Compared with other ecosystems in Wyoming, however, the grasshoppers of the Red Desert remain rather poorly known. The desert had never produced a full-fledged grasshopper outbreak, so the survey in this region was rather cursory.4 Less than 2 percent of the statewide sample was gathered from the Red Desert. But even this paucity of material suggests a remarkable biological diversity. The database reveals that the Red Desert supports 26 genera of grasshoppers, comprising 63 species. In other words, the 747 grasshoppers captured in the Red Desert represented 61 percent of Wyoming’s grasshopper biodiversity. Or consider that the grand total of grasshopper species in the United States is 630,5 so the Red Desert contains one in every 10 species found in the country. And there are reasons to suspect that undiscovered species are lurking in the Desert. Between Baggs and Point of Rocks, an entomologist specializing in grasshoppers has found what appears to be a new species in the genus Aeoloplides.6 Tim McNary has collected this creature at several locations, where it hides among the tangled branches of greasewood. The specimens visually differ from Wyoming’s only other species in this genus by having shorter wings and a smaller body, while also being notably darker than their tan cousins. If grasshoppers provide a representative picture of the insect fauna of the Red Desert, then the biological diversity is much higher than previously suspected. Indeed, there is the demonstrable potential for the discovery of entirely unknown life-forms in this ecosystem. However, the Socratic entomologist knows the depth of his ignorance. The reports of a few wise men or views

Jeffrey A. Lockwood

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 191

provided by omniscient databases cannot reveal the entomological riches of an ecosystem. For that, one must follow the Greek philosopher’s example and inquire for oneself. This is why my colleagues and I conducted what appears to be the first faunistic survey devoted solely to understanding the insects of the Red Desert. The first rule of field research is to take along the right people, which means collaborators who are more experienced, smarter, and stronger than oneself. The venture would have been a dismal failure had I not recruited Gary Beauvais, George Jones, and Scott Schell. Their geographical, biological, and pragmatic insights made it possible to take a meaningful ecological biopsy of the Red Desert. And “biopsy” is an appropriate term, as the venture could at best provide an infinitesimal sampling of a vast ecosystem.7 But such samples can, if carefully extracted, reveal important aspects of a person or a place. The second rule of insect faunistics is to take along supplies for as many different sampling methods as possible. Insects that might be captured by the hundreds using one technique could be completely overlooked by another. Our sampling equipment allowed six methods of collecting insects. We took along sweep nets, the staples of field entomology. This type of net consists of a canvas bag (able to survive being used amid gnarled shrubs) held open with a fifteen-inch metal hoop, which is mounted on a stout wooden handle. To collect ground-dwelling insects, we used pitfall traps, which is the technical term for sixteen-ounce plastic cups buried flush with the soil surface and charged with a few ounces of ethanol to kill and preserve whatever tumbles over the edge. The Wyoming version of the pitfall trap includes a plastic plate, supported on a tripod of metal legs, to form a cover about an inch above the trap. This additional feature tends to lure unsuspecting insects into a sheltered area, but the real reason for the cover is to slow the evaporation of the alcohol in dry, windy conditions. Flying insects are effectively collected with Malaise traps (Figure 8.1). These devices look like six-foot-tall tents made of net-

Figure 8.1. Malaise trap. Photo by Scott Schell.

ting; the “walls,” or baffles of the trap, intercept insects in flight. Once they land, the insects invariably crawl upward and into a collecting jar mounted at the peak of the trap. For sampling aphids and insects of their ilk, we used yellow sticky cards. Some herbivorous insects are attracted to bright yellow, which is why savvy rose gardeners often plant a single bush of a white or yellow variety to serve as a “trap crop” for plantsucking insects. In the field entomology version of this trapping system, the cards are coated in a tacky glue and mounted on wooden stakes. The result is a miniature, canary-colored version of La Brea tar pits. Nocturnal insects are attracted to lights, as anyone with a porch light can attest. The scientific adaptation of the backyard phenomenon consists of hanging a battery-powered black light next to a white sheet (Figure 8.2). In the right habitats, light traps can

Insects of the Red Desert

191

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 192

Figure 8.2. Insects drawn to a light trap. Photo by Scott Schell.

draw in hundreds of insects within minutes. Large specimens can be plucked from the sheet with forceps, and small insects can be gently aspirated using a gizmo that allows one to suck a specimen into a vial without inhaling the creature. Finally, we brought along a few pounds of oatmeal to set out in the evenings. A trail of oatmeal is the best way to attract one of the desert’s most remarkable denizens, the Jerusalem cricket — a bizarre subfamily of large, flightless crickets sometimes called “children of the earth” because their oversize round head with beady eyes is apparently reminiscent of a miniature, albeit grotesque, child. Actually, we didn’t know whether there were any Jerusalem crickets in the Red Desert, but dry, sandy habitats often harbor these insects. The third rule of a good survey is to sample from a diversity of habitats, including the dominant vegetation. We chose to set 192

up the Malaise, pitfall, and yellow sticky traps in a sagebrushgreasewood habitat that typified much of the Red Desert. The light trap and oatmeal trail were used in the Killpecker Sand Dunes. To allow direct comparisons among habitats, we took sweep net samples from sagebrush-greasewood communities, sand dunes (amid stands of rabbitbrush on the dunes and along sedges growing in a wetland formed at the base of the dunes), and an aspen grove incongruously flourishing on a high, cool ridge above the flats. And for final good measure we occasionally stopped to flip over roadkill and search for six-legged scavengers. Such a thoroughly planned expedition would have provided a reasonably compelling estimation of the Red Desert’s insect inhabitants, except for a single absurdity. A combination of commitments and logistics conspired to limit the window of opportunity for the team to a two-day period. With driving time factored into the venture, the entomological survey of nearly 24,000 square miles of desert was compressed into a twenty-fourhour period beginning with setting up traps in the greasewoodsagebrush flats at noon, August 29, 2005. Even without Socrates’ admonition, the idea of coming to know much of anything about insect life in such a vast and diverse region under these constraints would have been ridiculous, had it not been for a number of remarkable discoveries. The single day of sampling, representing about fifteen manhours of actual collecting, produced 3,364 specimens. Two dedicated student workers took more than 150 hours to pick the insects from sand, twigs, seeds, and other detritus and place the specimens into “look-alike” sets as a first approximation for subsequent classification. From this point, Kirk VanDyke sorted the insects into morphospecies—a term that entomologists use for taxonomic groupings that expert opinion would suggest are true species but for which no authoritative identification has been made. With nearly 100,000 species of insects in North America, no single entomologist can confidently provide a name for even 1 percent of the total fauna. Within the phylum Arthropoda, our samples represented 2 classes and 13 orders: Acari (mites) and

Jeffrey A. Lockwood

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 193

Araneae (spiders), from the class Arachnida; and Diptera (flies), Coleoptera (beetles), Hemiptera (bugs), Homoptera (aphids, plant hoppers, and their relatives), Hymenoptera (ants, bees, and wasps), Lepidoptera (butterflies and moths), Mantodea (mantids), Neuroptera (lacewings and their relatives), Odonata (dragonflies and damselflies), Orthoptera (grasshoppers and crickets), and Trichoptera (caddis flies), from the class Insecta. Considering that Fautin’s 1946 survey of arid shrublands yielded only 9 orders of invertebrates, perhaps a well-planned biopsy of the Red Desert wasn’t so absurd after all. Among the 11 orders of insects, we identified 83 families. This is a greater diversity in a single day’s sampling than the combined number of families of birds, mammals, fish, amphibians, and reptiles in the state of Wyoming (the birds are the most diverse vertebrates with 55 families, including all of the resident, migratory, and accidental species). The insect family most commonly encountered was Lygaeidae, the seed bugs, which were presumably taking advantage of the late-summer seed set in various plants. Thirteen of the families were represented by a single specimen each, so we might presume that there were many other rare families that our sampling methods overlooked. This higher-level taxonomy was quite impressive, but the biological diversity of species was truly remarkable. Our samples yielded 463 species of insects. Flies were the most speciose order, with 149 kinds. At the other end of the spectrum, we caught only a single species of caddis fly. Ecologists have various ways of expressing diversity, and the most intuitively appealing measure is Simpson’s D (named for the British statistician Edward Hugh Simpson, who proposed the index in 1949). For our sample, D was calculated to be 0.96. This value expresses a probability that can be understood with a simple thought experiment. Imagine that you have written down on separate slips of paper the name of every organism that was caught in a given habitat. Now you drop all these into a large bowl, mix the contents thoroughly, and select two pieces of paper. The probability that these slips have the names of different organisms written on them

is the value of Simpson’s D. So, for the Red Desert, the chances of randomly picking two insects and having them be different species is about 96 percent, or about nineteen chances in twenty. As you might guess, this is a very high level of diversity. Ecologists are well aware that any sample of organisms from a habitat cannot possibly include the entire gamut of species. Sweep netting during the day, for example, will miss grounddwelling and nocturnal insects. And even in the most thoroughly sampled habitat (greasewood-sagebrush), we used only a small portion of the available methods; there are techniques for extracting insects from plant litter and soil, for collecting parasites from their vertebrate hosts, for attracting dung-feeding insects, and so on. We know that we missed some entire taxa, because our sampling yielded no specimens of the order Thysanoptera (thrips) even though a survey by Tingey and his colleagues that targeted these tiny creatures found them to be relatively abundant in sagebrush-grass habitats. Given that any survey will overlook at least a portion of the organisms of interest, mathematical methods have been devised to estimate “sample coverage,” or the percentage of the total diversity that is represented by a particular study. This approach originated with ethologists who, after having spent long hours peering through binoculars in the field, wanted to know when they had seen most of an animal’s repertoire of behaviors. The underlying logic of the calculation is that the greater the proportion of kinds (behaviors, species, etc.) that have been observed only once, the greater the number that have not been seen at all. And in our sample, a rather large percentage (46 percent) of the 463 species were represented by just a single individual. Based on the “sample coverage” formula, we estimate that there were actually 861 species in our habitats. This suggests that there were nearly 400 kinds of insects—perhaps hiding in the soil, under stones, or inside plants or other insects—that our rapid assessment failed to find. But even this calculation represents an egregious underestimate of the richness of insect life in the Red Desert. Our projection of entomological diversity is based on samples

Insects of the Red Desert

193

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 194

taken within a twenty-four-hour window, and we know that many insects are collectible for only a brief period of time. Adults of these ephemeral species burst upon the scene for a few days or weeks, before leaving behind life stages that are nearly impossible to find. Many desert insects spend the vast majority of their lives as eggs, often so small and cryptic as to be nonexistent in terms of standard sampling techniques. Likewise, a botanist making a collection on a single day would not expect to find more than a small fraction of the plants in flower. A snapshot of insect life taken on an August day could not possibly capture more than one-third of the species present in a habitat, and even this would be quite a remarkable accomplishment. But if we make this assumption, then our estimate of insect diversity must be tripled, meaning that 2,583 species is a more sensible approximation. However, there is one more problem. We sampled only three habitats and, as will be seen, the overlap of species among these plant communities was extraordinarily limited. Our biopsy of the landscape completely missed the veritable forests of big sagebrush, the stands of juniper crowning ridge lines, and streams lined by riparian vegetation. And these are only a few of the more obvious habitats (bird nests, cliff faces, dung piles, deer hides, subsurface soils, and dead animals would have entirely different insect communities). Not even in the boldest moment of scientific arrogance could an ecologist claim that our sampling protocol captured more than half of the species found among the various habitats. Doubling our previous estimate puts us at 5,166 species of insects, and even this number is almost certainly far lower than the actual richness. In sum, the Red Desert must host more than 5,000 species of insects. Consider what such a number means in the context of other life-forms in Wyoming and the nation. According to the Wyoming Game and Fish Department, there are 657 vertebrate species in the state, including all native, exotic, resident, and accidental varieties of fish, amphibians, reptiles, birds, and mammals.8 We found, all estimates aside, 70 percent of this biological diversity among the insects collected in a single day. In total, the insect bio194

diversity of the Red Desert surely exceeds the number of vertebrate species in the entire United States (3,041). Even plant diversity pales in comparison with that of insects. While the number of plant species in the Red Desert is anyone’s guess, the entire state hosts 2,958 species, perhaps half as many as the number of insects in the Red Desert. How can such a harsh ecosystem give rise to such biological diversity? By understanding the differences in insect communities among the three sampled habitats, we can begin to formulate an answer—the Red Desert is an extraordinarily heterogeneous ecosystem. In the greasewood-sagebrush plant community, the most frequently collected insects were Lygaeidae, the second-largest family of Hemiptera in North America, with more than 250 species. The most familiar lygaeid, although not among the species in our samples, is the milkweed bug. While seed bugs were numerous, the family with the largest number of species (13) was an omnipresent fly, the Empididae. Few visitors to the Red Desert would suspect that these minute insects represent such diversity, but most people would notice the dance flies. It’s hard to ignore the swarms of gnatlike adults that give the family its common name by bobbing up and down in mass synchrony—and seem to relish doing so around the heads of humans. Although they are no risk to people, these flies are tiny predators, a feeding strategy that some species have parlayed into remarkable courtship rituals. Males neatly wrap fresh prey in silken packages to offer as nuptial gifts to prospective mates. And females must be highly discerning, for a male may try to cheat by offering only a bundle of silk wrapping to occupy the female while he attempts a quick copulation. The greasewood-sagebrush insect community differed from those of the other habitats in having much higher frequencies of five families. The grasshoppers (Acrididae) were far more abundant in this habitat. And our collected material included a specimen of Melanoplus bowditchi, the sagebrush grasshopper — a species not previously recorded from the Red Desert. As a family, grasshoppers are quite adept at partitioning resources, by spe-

Jeffrey A. Lockwood

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 195

cializing on particular soil types, plants (even plant parts), and other features of a habitat. In addition to these herbivores, a family of pollinators was strikingly abundant. The Halictidae, a group of ground-nesting communal bees, includes species that are painfully familiar to anyone who has spent time in arid lands. The sweat bees’ persistent efforts to lap up perspiration would be almost endearing if it were not for their propensity to deliver a burning sting when accidentally trapped beneath a shirt collar. We might rationalize that harsh environments foster violent tendencies, which would also account for the abundance of Therevidae. The stiletto flies are named for their tapered, knifelike bodies, but the allusion to a deadly weapon might just as well refer to their predaceous larvae, which feed on other arthropods in the sandy soil where they live. However, predators are only a part of the deadly story. The other two families of unique abundance in the greasewood-sagebrush habitat were parasites. The Braconidae is a group of wasps whose larvae, depending on the species (there are more than 1,700 in North America), feed inside the living bodies of a phenomenal range of insects and spiders. The last family distinguishing this habitat from the others in the Red Desert validates the entomological insight of Jonathan Swift: So, nat’ralists observe, a flea Hath smaller fleas that on him prey, And these have smaller still to bite ’em And so proceed ad infinitum. The Perilampidae are not fleas but hyperparasitic wasps—insects that parasitize other parasitic wasps. The perilampids lay their eggs on plant foliage, and the newly hatched larvae lie in wait for an unsuspecting caterpillar. The larval wasp bores into the caterpillar and sets up an ambush for another parasite, such as a braconid. Once the larger parasite has pupated, the hyperparasite penetrates the pupa, consumes its body, and finally

emerges from the well-used caterpillar. In the desert, where water and protein are in short supply, the competition for limited resources yields remarkable tales. Perhaps the strangest creature collected from the greasewoodsagebrush habitat was a single wasp from the family Dryinidae, very rare parasites that attack Homoptera. The wingless female resembles an ant with the front legs of mantid, which are used to hold onto the host—who is presumably not too happy about the situation— during oviposition. Imagine a Lilliputian version of a Wyoming bull rider and you get the picture. At least some species of dryinids are polyembryonic, meaning that a single egg gives rise to multiple offspring in a process similar to that which produces identical twins in humans (except that in the insect there are more than fifty identical siblings). But bizarre insects were not the exclusive purview of the greasewood-sagebrush flats. The insect community of the sand dunes had a plethora of Anthicidae (the most numerous insects in our collections), with the half-appropriate and long-winded common name “antlike flower beetles.” Although not particularly associated with flowers, these odd insects have the appearance and scampering gait of ants. The anthicids also have antlike feeding habits, being both scavengers and predators. The most diverse family in our samples was Ichneumonidae (14 species), a group of parasitic wasps with more than 3,100 species in North America. Like their sisters, the Braconidae, these wasps attack an impressive range of other insects, with various species specializing on particular hosts. Three families distinguished the sand dunes’ insect fauna from that of the other habitats. In addition to Anthicidae, Mycetophilidae and Byrrhidae were notably common. Mycetophilids are a familiar indoor nuisance—the veritable swarm of fungus gnats that aggregate in the vicinity of potted plants. The insects’ larvae flourish in damp soil with decaying vegetation, which describes both houseplants and low-lying areas at the base of sand dunes, where dead vegetation and moisture accumulate. Looking like miniature mosquitoes, these gnats also have achieved a measure of celebrity, at least by insect standards. The larvae of a bizarre,

Insects of the Red Desert

195

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 196

bioluminescent species of mycetophilids are the namesake of the famed glowworm caves of New Zealand. The pill beetles (Byrrhidae) are known for both their affinity for sandy habitats and their cowardice. When disturbed, they draw in their legs and play dead, resembling shiny, black BBs. The sand dunes also hosted a couple of creatures that were common in the aspen groves but virtually absent from the greasewood-sagebrush. The leafhoppers (Cicadellidae) are quarter-inchlong, bullet-shaped insects. Their color and morphology allow them to blend into the foliage as they suck sap from various plants. And the cicadellids are the host for the larvae of the strangest denizen of the dunes and aspen groves: Pipunculidae, a big-headed fly that looks like a hydrocephalic insect. The size of its head rivals that of the rest of its body, and the vast majority of the head is composed of a pair of monstrously large eyes. These uncommon flies presumably use an acute sense of vision to pick out their cryptic victims from amid the vegetation. Despite these similarities between the insect communities of the sand dunes and the aspen grove, the latter habitat also had its own distinct insect fauna. The most numerous insects collected from the aspen grove were Sepsidae, a family of flies whose larvae scavenge in rotting vegetation and animal excrement. Another group of flies, the Anthomyzidae, was the most diverse family, with 5 species in our samples. This might not seem like a great deal of diversity, but there are only 14 species of anthomyzids in all of North America, and finding more than one-third of any insect family’s species in a single habitat is rather remarkable. These flies are often common in grassy and marshy areas, which may account for why they flourished in the relatively lush oasis of the aspens. Along with Sepsidae, the family that distinguished this treed sanctuary from the other habitats was Aphidae. The aphids, or plant lice, are soft-bodied insects familiar to gardeners worldwide and the nemesis of Wyoming homeowners with cottonwood trees that rain honeydew—the sugary liquid excrement of aphids— onto cars, porches, and sidewalks throughout the summer. These 196

insects exhibit a phenomenal range of life histories, including the capacity for “telescoping generations” in which a female aphid has within her body an offspring (some species give live birth) that has within its body an offspring. A female carrying her own granddaughter is made possible by parthenogenesis, a form of reproduction in which females effectively clone themselves without mating. Given the fleeting and unpredictable flush of greenness within an aspen grove surrounded by desert, insects must take advantage of ephemeral resources. And with their reproductive capacity, the aphids are biologically poised to rapidly exploit such ecological windows of opportunity. The clearest picture of the differences in the insect communities among the habitats of the Red Desert arises by comparing the samples from sweep netting, a uniform method of collecting used at all sites. Although many comparisons can be made, a few examples will suffice to illustrate the entomological uniqueness of the habitats. Fewer than one in five of the insect families found in the greasewood-sagebrush community (21 families) were also found in either the sand dunes (52 families) or the aspen grove (37 families). The differences were even more striking in comparisons of species. Of the 32 species from the greasewood-sagebrush and the 172 species from the sand dunes, just 3 species were collected in both habitats. Likewise, there were just 3 species in common between the greasewood-sagebrush and the aspen grove. The pattern of uniqueness among insect communities also was evident at much finer scales than that of the three major habitat types. The sweep net samples taken from the drier microhabitats on top of the dunes and those collected from the moister areas along the base of the dunes were as different as the insect communities from greasewood-sagebrush flats and aspen groves. Of the 125 species collected from the xeric slopes and the 48 species collected from the hydric areas of the dunes, only two species were found in both microhabitats. For insects, the rabbitbrush flourishing on the upper region of a sand dune is a world utterly unlike the sedges growing on the wet sands a hundred yards away.

Jeffrey A. Lockwood

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 197

Even along a fifty-yard transect, we found extraordinary partitioning of the habitat, as evidenced by different sampling methods. In the greasewood-sagebrush habitat, just 2 of the 80 species that fell into the pitfall traps were also among the 33 species caught in the Malaise traps. Even more remarkably, not a single species caught by the Malaise traps was also found among the 35 species collected with sweep nets. The insect world on the surface of the soil was virtually unknown to that of the air, which in turn had nothing in common with that of the foliage. Our ecological biopsy also made it clear that insects of the Red Desert divide up resources not only in space but also in time. Light trapping and sweep netting revealed insect communities as different as, well, night and day. In the sand dunes, only 4 of the 63 species attracted to the light trap were among the 127 species found in the sweep net collections. In the name of humility and honesty, I should also admit that not all forms of trapping were successful in adding pieces to the faunal puzzle of the Red Desert. Our oatmeal trails were a culinary boon for rodents, whose nocturnal larceny was abundantly clear from footprints in the sand. No Jerusalem crickets came to the banquet, or perhaps the rodents converted the lumbering insects into dessert. And the next morning, howling winds coated the yellow sticky cards with a fine grit, effectively converting them into fluorescent sandpaper and making them safe for insects. In the end, of course, no amount of collecting, counting, and calculating allows us to truly understand a place as starkly beautiful as the Red Desert. But at least some of the qualities of this paradoxically rich and austere landscape can be inferred from the samples that we extracted. First, the insects of this ecosystem have evolved remarkable degrees of specialization. There appear to be very few generalists, species capable of occupying more than a single habitat. Even within a habitat, we see evidence of hyperspecialization, with instances in which more than a dozen species in a single family partition the available resources. And within guilds, such as the parasites, there are ecological hierarchies, with parasites beset by

hyperparasites. The enormously complex food webs, which appear to be dominated by insects, squeeze every bit of water and nutrients out of the habitat. Such ecological elaboration may well be the consequence of life in a cold desert, where conditions are harsh but highly predictable. Of 144 weather stations in Wyoming, two of the three reporting the least variation in precipitation over the last half century are in the Red Desert (Bitter Creek and Wamsutter).9 Predictability may be as important as productivity in fostering the evolution of biological diversity. Next, the extreme endemism within the Red Desert is cause for grave concern with respect to anthropogenic disturbances, even those which are highly localized. Once species are lost from secluded communities, the rate of recolonization will be extremely slow. Moreover, the intimate linkages among species may require a very particular sequence or timing of fortuitous arrivals, further delaying or even preventing natural restoration of a damaged habitat. The standard approach to restoration ecology might accelerate recovery. That is, if representative instances of intact habitats are preserved as “biotic banks,” these areas could function as sources for sustainable withdrawal of specimens in the reclamation of disturbed sites. However, this classic model of restoration ecology may not be viable in the Red Desert. We do not know whether all or most instances of a particular habitat (e.g., the isolated wetlands at the bases of sand dunes or spatially disjunct greasewood-sagebrush flats) harbor the same communities. It may well be that habitat islands with similar plant communities support unique insect faunas, and we have good, albeit limited, reason to suspect that such is the case. Without more systematic and detailed entomological surveys, it is impossible to know whether the standard methods of ecological restoration are applicable to this ecosystem. Finally, given that no more than one in ten of the world’s species have been described10 and that no systematic survey of the insect fauna of the Red Desert has been undertaken, it seems reasonable to suggest that there are potentially hundreds, and surely dozens, of species unknown to science in this unexplored

Insects of the Red Desert

197

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 198

ecosystem. And in light of the ongoing and proposed coal-bed methane extraction projects, there is a very high likelihood that intense, localized disturbances will lead to extinctions. We will, in effect, destroy marvels of the natural world—life-forms that struggled for millennia to carve niches in the vastness of the desert— without ever knowing they existed. In the final analysis, based on our absurdly limited information about the Red Desert, only one conclusion seems evident. We have an abundance of ignorance and need a wealth of humility. For the Socratic entomologist trying to understand the nature of the Red Desert and the risks that human activity might have on this paradoxical place with its riches of life amid a poverty of water and nutrients, the modern-day insights of Wendell Berry are fiercely apropos: “The real question that is always to be addressed is the one that arises from our state of ignorance: How does one act well—sensitively, compassionately, without irreparable damage— on the basis of partial knowledge?”

198

Notes 1. I had to add “Wyoming” to this search to eliminate sites pertaining to Australia, which has its own Red Desert. 2. Anonymous, ed., The Sagebrush Ecosystem: A Symposium (Utah State University, Logan, 1979), 56. R. W. Fautin, Biotic communities of the northern desert shrub biome in western Utah, Ecological Monographs 16 (1946): 251–310. G. B. Hewitt, E. W. Huddleston, R. J. Lavigne, D. N. Ueckert, and J. G. Watts, Rangeland Entomology (Society for Range Management, Denver, 1974). C. D. Jorgensen and W. F. Tingey, Insects of the Sagebrush-Grass Community in Utah, Program Report of the Intermountain Forest and Range Experiment Station 12-11-204-31 (1968). W. M. Tingey, C. D. Jorgensen, and N. C. Frischknecht, Thrips of the sagebrush-grass range community in west-central Utah, Journal of Range Management 25 (1972): 304–308. 3. The Wyoming Grasshopper Information System can be accessed through the “Grasshoppers of Wyoming and the West” website, http://www.sdvc.uwyo.edu/grasshopper/ghwywfrm.htm. 4. Greater than 90 percent of the Red Desert has never had a grasshopper outbreak, defined as a density of at least eight grasshoppers per square yard. Such outbreak densities have been recorded only along the northern edge of the desert during major infestations of the eastern plains. However, during the expedition that my collaborators and I made in 2005, we found several grassy openings amid the sagebrush near the Boar’s Tooth that supported ten to fifteen grasshoppers per square yard. So either we were incredibly fortunate to find the first grasshopper outbreak in history that reached the heart of the Red Desert or nobody has devoted much energy to surveying this ecosystem. We suspect the latter. 5. J. L. Capinera, R. D. Scott, and T. J. Walker, Field Guide to Grasshoppers, Katydids, and Crickets of the United States (Cornell University Press, Ithaca, N.Y., 2004). 6. J. A. Lockwood, T. J. McNary, J. C. Larsen, and J. Cole, Distribution Atlas for Grasshoppers and the Mormon Cricket in Wyoming, 1988 – 1992, University of Wyoming Agricultural Experiment Station Bulletin B-976 (1993). This document is also available through the “Grasshoppers of Wyoming and the West” website, http://www.sdvc.uwyo.edu/ grasshopper/ghwywfrm.htm. 7. In the course of our work, we sampled perhaps 15 of the 15 million acres that make up the Red Desert. Likewise, a tenth of a gram of tissue biopsied from a 220-pound adult would represent a millionth of the total.

Jeffrey A. Lockwood

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 199

8. Official lists of the vertebrates occurring in the state are available at the Wyoming Game and Fish Department website, http://gf.state. wy.us. 9. B. E. Martner, Wyoming Climate Atlas (University of Nebraska Press, Lincoln, 1986). 10. E. O. Wilson, ed. Biodiversity (National Academy Press, Washington, D.C., 1988).

Insects of the Red Desert

199

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 200

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 201

9. Sagebrush George P. Jones

What Is Sagebrush?

chemistry

Botanists share an understanding that “sagebrush” means a group of some dozen species or subspecies of shrubs of western North America, part of the large, worldwide genus Artemisia. Whether the woody progenitor of sagebrush evolved in Eurasia and migrated to North America or arose in North America from a more distant herbaceous ancestor is not known. (Recent interpretations favor the latter.) Two things are clear: the present group of species radiated from a common ancestor in North America during alternating periods of moist and dry climates that repeatedly isolated populations from one another; and sagebrush was a prominent part of the western North American flora by the late Tertiary period or early Quaternary period, 1.6 to 2 million years ago.1

Every type of sagebrush is aromatic (as are many other species of Artemisia) due to volatile chemicals in the leaves and flowers. Different sagebrush species and even different subspecies have recognizably different aromas. Whether or not the aromatic chemicals of sagebrush evolved as a deterrent to browsers in the endless battle between the consumed and the consumers, they do seem to influence the animals that eat sagebrush. Greater sagegrouse enclosed in an area with several sagebrush types first consumed the type with the least amount of volatile compounds, then the type with the next-lowest amount, and so on.2 More recent research with mule deer on the northern range in Yellowstone National Park has shown that chemical compounds extracted from subspecies of big sagebrush (Artemisia tridentata) and black sagebrush (Artemisia nova), applied to chopped alfalfa feed, reduced the intake of that feed.3 When browsing on sagebrush on the northern range, mule deer preferred the big sagebrush subspecies with the lowest chemical concentrations and ate little of the black sagebrush, which was highest in those compounds.4 Concentrations of some chemicals have been shown to change in the same plant from one season to the next and even between morning and evening of the same day.

Characteristics of Sagebrush The remarkable success of sagebrush taxa throughout the west is the result of a set of shared characteristics. (See Table 9.1 for a list of the species and subspecies of sagebrush.)

201

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 202

Table 9.1. Occurrence of sagebrush species and subspecies in the Red Desert Scientific name1

Common name2

In the Red Desert?3

Artemisia arbuscula ssp. arbuscula Artemisia arbuscula ssp. longiloba Artemisia bigelovii Artemisia cana ssp. cana Artemisia cana ssp. viscidula Artemisia nova Artemisia pygmaea Artemisia rigida Artemisia rothrockii Artemisia tridentata ssp. spiciformis

Little sagebrush Alkali sagebrush Bigelow sagebrush Silver sagebrush (basin silver sagebrush) Silver sagebrush (mountain silver sagebrush) Black sagebrush Pygmy sagebrush Scabland sagebrush Timberline sagebrush Big sagebrush (subalpine big sagebrush)

Artemisia tridentata ssp. tridentata Artemisia tridentata ssp. vaseyana Artemisia tridentata ssp. wyomingensis Artemisia tripartita ssp. rupicola Artemisia tripartita ssp. tripartita

Basin big sagebrush Mountain big sagebrush Wyoming big sagebrush Wyoming three-tip sagebrush Three-tip sagebrush

Absent (western North America, including Wyoming) Minor, peripheral Absent (southwestern USA and Texas) Minor, peripheral Minor Minor Absent (southwestern USA) Absent (Pacific northwest) Absent (California) Absent (High-elevation Rocky Mountains, including Wyoming) Present Present Abundant Minor, peripheral Absent (northwestern USA, including Wyoming)

Note: The eleven taxa in this table are those listed in E. Durant McArthur and A. Perry Plummer, “Biogeography and Management of Native Western Shrubs: A Case Study, Section Tridentatae of Artemisia,” in Intermountain Biogeography: A Symposium, Great Basin Naturalist Memoirs, no. 2 (1978). 1. usda, nrcs, The plants Database, http://plants.usda.gov (National Plant Data Center, Baton Rouge, La., 2005). 2. Where two common names are given, the first is from the standardized list of plant names for the USA (ibid.) and the second (in parentheses) is a name in common use in Wyoming. 3. Whether a sagebrush taxon is present in or absent from the Red Desert is noted here. For taxa designated as absent, the geographic distribution is noted briefly.

Despite the chemicals, sagebrush is an important food for wildlife. Julius Nagy, in his review of research on the topic, pointed out that sagebrush has relatively high nutritional value in late winter, when other potential food plants are unavailable or low in nutrients.5 He noted that late winter may be the time of year when concentrations of volatile chemicals are at their lowest in the sagebrush. Discussions of animals and sagebrush usually concentrate on vertebrates, but in most ecosystems insects eat far more vegetation. The possibility that volatile chemicals are present in sage202

brush to discourage feeding by insects has received scant attention from researchers. An altogether different hypothesis is that the volatile chemicals suppress plants that otherwise would compete with sagebrush for water or nutrients. Extracts of chemicals from sagebrush leaves have been shown to reduce germination of grass seeds in experimental settings. But as Dennis Knight noted in his comprehensive review, this falls short of demonstrating that the volatile compounds improve the competitive status of sagebrush in the field.6

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 203

leaves and water loss All sagebrush species except silver sagebrush (Artemisia cana) are evergreen. Unlike evergreen coniferous trees, which keep each leaf for several years, sagebrush annually produce two crops of leaves that stay on the plant for a year at most. A sagebrush plant comes out of winter bearing small, overwintering leaves. In early spring the plant produces new, relatively large ephemeral leaves, and soon thereafter the overwintering leaves begin to drop. After the ephemeral leaves have grown, a new crop of overwintering leaves is produced in bunches at the bases of the ephemeral leaves, replacing the former year’s overwintering leaves and staying on the plant until the following spring. The ephemeral leaves start to fade and drop from the plant in midsummer; the new overwintering leaves remain, allowing the sagebrush the ability to photosynthesize in favorable conditions throughout the year. Both ephemeral and overwintering leaves on the vegetative branches (that is, the branches without flowers) of all species except silver sagebrush are wedge-shaped and attached to the branches at the narrow end. At the broader, distal end, they are divided into three lobes. The overwintering leaves are a primary characteristic used to identify plants to species or subspecies, because the sizes of these leaves, their width-to-length ratios, and the shapes of their lobes differ among taxa. (Ephemeral leaves are highly variable in size and shape.) The scientific names of two common Wyoming species are taken from the shapes of the leaf lobes: on Artemisia tridentata, big sagebrush, the three lobes extend only a short way back from the leaf tip and resemble three small teeth, while on Artemisia tripartita, three-tip sagebrush, the leaves are cut deeper into three obvious parts. Within the widespread and abundant big sagebrush, the overwintering leaves help to distinguish subspecies. The tall, shaggy basin big sagebrush (Artemisia tridentata ssp. tridentata) has long, narrow leaves with very shallow teeth.7 Mountain big sagebrush (Artemisia tridentata ssp. vaseyana) has broader leaves with larger teeth. On Wyoming big sagebrush (Artemisia tridentata ssp. wyomingensis), the leaves

are short and broad, with especially small teeth. Leaves are the sites of photosynthesis, the fundamental process through which carbon dioxide (CO2) is converted into the compounds that make up the plant body. They are also the sites of a serious and unavoidable trade-off: a plant cannot absorb CO2 from the atmosphere without losing water to the atmosphere, and the faster the rate of CO2 uptake and photosynthesis, the faster the rate of water loss. The large ephemeral leaves that a sagebrush bears in spring and early summer provide the plant with its maximum amount of photosynthetic surface while soil moisture is plentiful. That is the time of year when sagebrush grows the fastest, as the new overwintering leaves are produced and stems and roots elongate. The previous year’s overwintering leaves are, by then, unproductive, and maintaining them would result in a net loss of food and a waste of water. So from the point of view of an energy and water balance, the sagebrush is better off shedding the old overwintering leaves and putting its resources into the ephemeral leaves. The graminoids and forbs also are growing rapidly and using water during spring and early summer.8 Gradually, the soil water supply is depleted as plants remove water and as the rising summer temperatures cause more water to evaporate from the soil surface. Water stress becomes more severe in plants as the warmer air causes more water to be lost from leaves while water becomes harder for the roots to remove from the soil. The ephemeral leaves of sagebrush, with their relatively high rates of food production and water loss, are shed. In their place are the smaller overwintering leaves. Rates of water loss and of growth decline, but the sagebrush continue to grow slowly and produce flowers and, eventually, seeds. Most of the shallower-rooted grasses and forbs, in contrast, become dormant; they have completed their growth for the year. By maintaining small leaves, sagebrush is able to prolong its growing season well into the fall and begin growth early in the spring. Sagebrush have even been shown to photosynthesize on warm winter days where the soil is unfrozen.

Sagebrush

203

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 204

Silver sagebrush produces only one set of leaves and is not evergreen. Its leaves usually are unlobed but may be divided into uneven lobes that look quite different from the teeth of the other sagebrush species. These leaves appear in the spring and drop in the fall and early winter. The reason for this deciduous habit, unusual among the sagebrushes, is unclear. Silver sagebrush grows along streams and in bottoms of draws where soils remain moist throughout the growing season, and consequently these plants probably avoid the degree of water stress experienced by sagebrush on drier sites. Leaves of most sagebrush species are covered with short silver hairs that nearly obscure the leaf epidermis. Light-colored hairs in some plant species have been shown to reduce leaf temperature and water loss in sunny habitats and appear to be an adaptation for saving water, and this may be true for sagebrush too. Hairs also can discourage insect herbivores, thereby protecting the plant’s investment of energy and nutrients in leaves. If you gently scrape away the hairs and examine the leaf with your hand lens, you can see the darker epidermis with its tiny, dotlike glands. Black sagebrush, as its name suggests, lacks silver leaf hairs. You can see this, too, through your hand lens. This taxon’s dark color is one characteristic that distinguishes it from the gray Wyoming big sagebrush with which it often grows. roots In arid lands such as the intermountain basins, competition for water is stronger than competition for light, and the greater part of a plant grows below the ground. The size and extent of a plant’s roots depend greatly on the rockiness and depth of the soil, but researchers at several sites around the west have found that sagebrush generally has a root system that extends deep into the soil and spreads widely around the plant. Hydrologist David Sturges documented the relative sizes of the aboveground and belowground sagebrush plant.9 On plants excavated at a research site west of Saratoga, Wyoming, the rooting depth ranged from forty204

eight inches in the shallower soil along the upper slope to over sixty-six inches in the deep soil at the bottom of the slope. Heights of the shrub canopies, in contrast, averaged only eight and onehalf inches at the upper slope and twenty-four inches at the bottom. Roots extended outward forty-eight to sixty inches from the plant, with the most widely spreading roots in the upper twelve inches of the soil. Despite the extent and depth of the root system, though, most of the roots, particularly the small roots most active in absorbing water and nutrients, were close to the plant and near the soil surface. In western North America, the soil water reservoir usually is recharged to its annual maximum in spring by melting snow and rain. Measurements of soil water around sagebrush plants at the Saratoga site showed that they began using water from the upper part of the soil in spring, and withdrew water from greater distance and depth as the season advanced. Uptake of water and nutrients by sagebrush roots is augmented by beneficial associations that the shrubs form with soil fungi. These associations, called mycorrhizae, are common among shrubs, graminoids, and forbs of arid lands. In sagebrush, they form when threads of the fungus grow into the smallest of the plant’s roots. The plant in a mycorrhizal association benefits because the fungal strands growing throughout the soil greatly increase the surface area through which water and nutrients are absorbed. The fungus benefits by receiving energy that its host has produced through photosynthesis. flowers and seeds Midsummer is the start of the flowering season for sagebrush (except alkali sagebrush, Artemisia arbuscula ssp. longiloba, which flowers in early summer to midsummer). Sagebrush flowers are tiny (about one-eighth inch long), the petals fused into an inconspicuous, five-part corolla. Several flowers are borne together in a head characteristic of the sunflower family. In favorable years, thousands of heads on the upper part of the plant overtop the leafy

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 205

canopy. During late summer, the pollened mass gives the sagebrush a yellow tinge. With a myriad of flowers at the top of the canopy, sagebrush are good examples of plants depending on wind for pollination. The number of flowers on a sagebrush indicates the amount of soil water available to the plant. In many years, only the plants growing in draws or on lee slopes where snow accumulates sport thick flowering branches, while the plants on drier sites bear only a few heads or even fail to flower at all. Sagebrush seeds are as small as one would expect from the diminutive flowers, about one-sixteenth inch long, olive green, and roughly triangular. They mature late in the fall and drop to the ground near the plant. Most that germinate do so the following spring when soils are wet and warm. Seeds that do not germinate in the first year may remain viable for several years. seedlings Sagebrush plants add stem growth rings annually, so the age and year of establishment of a sagebrush plant can be determined from a count of growth rings in a cross section of the stem. Barry Perryman and other researchers at the University of Wyoming and Colorado State University recently reported on a study using this approach in stands of big sagebrush from around Wyoming.10 Their data showed a pattern known to population ecologists as “pulsed recruitment.” In a few years there had been pulses of abundant seedling establishment; in other years only a few seedlings had become established, and in most years, none. The failure of seedlings to survive in the first few years no doubt plays a larger role in determining the distribution and abundance of sagebrush than does survival of established adult plants. The tiny sagebrush seeds germinate into small seedlings whose roots reach only an inch or two into the soil for the first weeks. The surface soil dries out quickly in the warm sun, and shallowrooted seedlings die unless rainstorms come often enough to keep the upper soil moist and cool. The episodes of sagebrush seedling

establishment documented by Perryman and his co-workers in Wyoming have not yet been compared with weather records, but in British Columbia similar research showed that big sagebrush seedlings had become established in years with cool, relatively wet spring weather.11 Field experiments have shown that established stands of grass suppress sagebrush seedlings. Competition for water is a likely mechanism. This is the situation that sagebrush seedlings face after fires, which usually kill the sagebrush but not the grasses. In contrast, research in reclamation of surface mines has shown that sagebrush seedlings can become established in those highly disturbed environments (where all plants are starting out from seed) even in the face of competition from new grass seedlings. This research has also documented the importance of the early formation of mycorrhizal associations. Seedlings with mycorrhizal fungi can tolerate slightly drier soils than can seedlings without the fungi, and so can survive a few days longer between rainstorms. The Wyoming Big Sagebrush Steppe Sagebrush steppe covers most of the Red Desert. In the mid1990s, researchers from the Wyoming Gap Analysis Project used satellite images to map land cover throughout Wyoming,12 and their Wyoming big sagebrush cover type is the primary vegetation on 71 percent of the Red Desert (as defined here). Naturalists and writers who refer to the Red Desert as a sagebrush sea are concentrating on the Wyoming big sagebrush steppe that blankets the expanses of rolling hills and broad draws. Wyoming big sagebrush is the dominant shrub, but even brief observation shows that it virtually always grows with other shrubs. Douglas rabbitbrush (Chrysothamnus viscidiflorus) is common nearly everywhere. Rubber rabbitbrush (Ericameria nauseosa), gray horsebrush (Tetradymia canescens), and spiny hopsage (Grayia spinosa) are widespread but less common. The ubiquity of these shrubs on the landscape is revealed when they flower. By midsummer, spiny hopsage bears subtle red bracts on its flowering stems and horsebrush sports bright

Sagebrush

205

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 206

yellow flower heads. In late summer, the brilliant yellow heads of the rabbitbrushes illuminate the sagebrush steppe. The common name “big sagebrush” is misleading because, nearly everywhere in this vegetation, the shrubs are only kneehigh or shorter and form an open layer easy to walk through. The common name was originally applied to basin big sagebrush (Artemisia tridentata ssp. tridentata), which typically is larger than other sagebrush species. The smaller Wyoming big sagebrush (Artemisia tridentata ssp. wyomingensis) inherited the name. The herbaceous component of the sagebrush steppe growing on sandy and loamy sand soils is dominated by two bunchgrasses, needle-and-thread grass (Hesperostipa comata) and Indian ricegrass (Achnatherum hymenoides). Easily overlooked is a smaller bunchgrass, Sandberg bluegrass (Poa secunda), present in shrub steppe and grasslands throughout the west and noteworthy because it starts growing early in the spring. This hardy little native can even, in warmer sites, grow in the fall and keep its green leaves over the winter. Rhizomatous or sod-forming grasses are relatively minor components of the vegetation on these soils, but one species, thickspike wheatgrass (Elymus lanceolatus ssp. lanceolatus), is present in many places. Forbs grow everywhere but nearly always in low numbers, and only one, hoary goldenweed (Machaeranthera canescens), seems to be a regular member of the vegetation. The succulent pads of plains prickly-pear cactus (Opuntia polyacantha) are easily found but rarely are they dense enough to bother the poorly shod visitor. Where the soils are finer-textured in the sagebrush steppe, most of the same species are present but the dominance of needle-and-thread and Indian ricegrass is weaker, and Sandberg bluegrass and thickspike wheatgrass are more common. Black greasewood (Sarcobatus vermiculatus), shadscale saltbush (Atriplex confertifolia), and the low-growing Gardner saltbush (Atriplex gardneri) may join the other shrubs. Among the plants of the sagebrush steppe, only sagebrush is evergreen. The grasses and forbs complete their growth before the shallower soil water is gone and are dormant in most years by 206

late summer. The other shrubs are cold-deciduous and so drop their leaves by early winter. This seasonal difference in activity of the plants explains the shifts in diets of three animals closely identified with the sagebrush steppe. Pronghorn eat sagebrush yearround in the Red Desert, and studies of their food habits have shown that it constitutes from 70 –95 percent of their diet in winter.13 In spring and summer, when other shrubs, grasses, and forbs are growing, pronghorn eat less sagebrush and more forbs. Rubber rabbitbrush also is an important food then. Greater sagegrouse and pygmy rabbits, too, eat a high proportion of forbs in the spring and early summer, but little other than sagebrush in the fall and winter.14 In contrast, much of the research on mule deer (conducted elsewhere than the Red Desert) suggests that sagebrush is less important a food for them than it is for pronghorn, greater sagegrouse, and pygmy rabbits. Still, data collected near Saratoga, Wyoming, showed that sagebrush was the most important single item of food in the winter diets of mule deer, constituting 10 –55 percent of the food consumed.15 No doubt the mule deer that winter in the Red Desert also depend to some extent on sagebrush. Variation within the Wyoming Big Sagebrush Steppe A traveler through the Red Desert may notice patterns in the vegetation that reveal much about the environment of the big sagebrush steppe.16 The most striking may be those caused by fire. Wyoming big sagebrush is easily killed by fire and must reestablish itself from seed. The other shrubs in the Wyoming big sagebrush steppe, though, sprout from stumps after fire, and the grasses and forbs generally survive and sprout from root crowns or rhizomes. Hence a fire in the sagebrush steppe produces, for the first year or two, grass vegetation, then a mixed-shrub steppe as the sprouting shrubs grow up. Burned sagebrush stumps are a telltale sign of the cause of this unusual vegetation. Fire history is far more difficult to document in sagebrush steppe than in forest, and there seems to have been no effort to

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 207

do so in the Red Desert. Researchers working in big sagebrush vegetation in the Intermountain region, though, have concluded that, before settlement by Europeans, a typical dry sagebrush site burned every 35–100 years — a very broad range.17 Common sense suggests that dry habitats should be especially prone to fire, but that is true only to a point. Aridity limits the amount of plant growth and thus the amount of fuel, so dry habitats may burn infrequently. Because much of the big sagebrush steppe in the Red Desert produces little fuel, fires may well have been infrequent there and so have had relatively little effect on the pattern of sagebrush. More subtle and widespread than burned patches are clumps of relatively tall shrubs scattered throughout the shrub steppe. These sagebrush islands, as they are known among ecologists and naturalists, cover several hundred to several thousand square feet and stand a foot or two above the surrounding vegetation. The predominant shrubs are robust individuals of the same species that grow in the surrounding shrub steppe (Wyoming big sagebrush, Douglas and rubber rabbitbrush, spiny hopsage, black greasewood), and often they are joined by a tall native grass, basin wildrye (Elymus cinereus). Some of the sagebrush islands obviously grow on small dunes of silt or sand that provide a reservoir of soil water. The dunes may have formed in a time of drier climate or heavier grazing, when drifts of soil formed downwind from tall or dense shrubs and then expanded as more shrubs grew in the favorable substrate. In other places, though, the sagebrush islands do not sit atop obvious dunes, and their origin is a mystery. Ground squirrel and badger burrows are common in them, and perhaps those animals cause the islands by loosening the soil, allowing deeper penetration of soil water and shrub roots. An island may have started with a few tall shrubs growing around a single burrow and expanded as more mammals disturbed more soil around the tall shrubs that hid them from hawks. Throughout the Wyoming big sagebrush steppe, topography has an enormous influence on vegetation patterns. In the Red Desert, topography exerts its greatest influence by controlling the

deposition of snow. And snow is the crucial source of soil water. Deep roots supply the soil water that supports sagebrush late in the growing season. David Sturges documented the yield of water from snow in the sagebrush steppe west of Saratoga.18 By early April the snow that had accumulated over the winter midway down a lee slope contained three and a half times the amount of water as did the snow on the upper slope, and the snow at the foot of the slope contained five times the amount of water found at the upper slope. Snow was by far the greatest source of soil water; summer rains wetted only the upper part of the soil. Throughout the Red Desert, the consequence of increased soil water from drifted snow is seen in the dense stands of taller shrubs growing in draws and on lee slopes and, after midsummer, in the contrast of the thick flowering branches on the plants there with the sparse inflorescences on plants elsewhere. In especially dry years, the sagebrush plants on windblown slopes do not flower at all, and flowering branches are seen only on the shrubs that have tapped into the deep soil water provided by melting snowdrifts. Sagebrush not only responds to patterns of snow distribution but also helps create them. Any object that slows the wind below a threshold speed causes the snow to drop out of the air. In shrub steppe, topography largely controls the redistribution of snow by the wind: snow is removed from windward slopes and from flats where wind velocity is greatest, and then is deposited in draws and on leeward slopes where the wind slows down. But vegetation, too, slows the wind near the ground and captures some of the snow. The amount of snow stored in this manner is limited by the height and thickness of the vegetation, so sparse grass and dwarf shrubs store less snow than does Wyoming big sagebrush. Other Sagebrushes in the Shrub Steppe Wyoming big sagebrush dominates the sagebrush steppe throughout the Red Desert, but three other short sagebrush species grow with it in a few locales. Of these, the most wide-

Sagebrush

207

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 208

spread and common is black sagebrush. Aven Nelson, Wyoming’s preeminent botanist, described it as a separate species in 1900 and gave it the scientific name Artemisia nova, the “new artemisia.” Black sagebrush are nearly always short, rarely exceeding twelve inches. This species is said to be associated with calcium carbonate–rich geologic substrates, and in Wyoming it is found mainly in the foothills of mountain ranges near outcrops of limestone or sandstone. Usually it grows with Wyoming big sagebrush, but it can be found in pure stands covering several acres. Only a few areas in the Red Desert support black sagebrush. The Rawlins Uplift is one such place. At the southern foot of the Ferris Mountains farther north and in the Antelope Hills immediately south of the Sweetwater River, black sagebrush is common in the sagebrush steppe and dominates in some areas.19 A few plants grow with Wyoming big sagebrush at Cyclone Rim on the northern edge of the Great Divide basin. Studies of wildlife diets in the Red Desert have featured big sagebrush, not black sagebrush, and limited information on wildlife diets in other deserts is available. Pronghorn in a study in western Utah relied on black sagebrush throughout the year in the same manner that Red Desert pronghorn rely on Wyoming big sagebrush,20 but black sagebrush is the only sagebrush on that Utah range, so the study says nothing about its relative palatability or importance as food. At least we can infer that pronghorn eat black sagebrush where they find it in the Red Desert, and that it might be an important food source where it is common in the vegetation. Mule deer consumption of black sagebrush has received more study (but, as with pronghorn, not in the Red Desert). Researchers concluded that deer on Yellowstone’s northern range prefer big sagebrush21 and that some populations of black sagebrush are more palatable than others.22 We can only conclude that mule deer may eat black sagebrush in the Red Desert when it is available. The second of the low sagebrush species in the Wyoming big

208

sagebrush steppe is alkali sagebrush (Artemisia arbuscula ssp. longiloba), which grows with Wyoming big sagebrush, black sagebrush, and Wyoming three-tip sagebrush (Artemisia tripartita ssp. rupicola, discussed below) in the Antelope Hills south of the Sweetwater River. That vegetation is a mosaic of mixed stands where two, three, or all four species grow together, as well as pure stands of only one species. Plants of alkali sagebrush are difficult to distinguish from Wyoming big sagebrush. They are usually smaller, though, and their leaves are a slightly darker gray. Alkali sagebrush is unusual among sagebrushes in producing relatively large flowering heads early in the summer. The other species all flower later, and their flowering heads are noticeably smaller. This difference allows alkali sagebrush to be distinguished from Wyoming big sage for much of the year. The third dwarf sagebrush, Wyoming three-tip sagebrush (Artemisia tripartita ssp. rupicola), is the most unusual. This is truly a Wyoming sagebrush, found (so far) only in this state. It is common in shrub steppe on the foothills and lower slopes of the central and southeastern mountains. In the Red Desert this diminutive sagebrush, rarely taller than six inches even in flower, grows with Wyoming big sagebrush and black sagebrush at the southern foot of the Ferris Mountains, and with Wyoming big sagebrush, black sagebrush, and alkali sagebrush in the Antelope Hills. The granitic and limestone bedrock on which Wyoming three-tip sagebrush is found is missing elsewhere in the Red Desert, so the plant may be limited to those sites on the northern margin. The three deep lobes on its leaf ends make this sagebrush distinctive, and it is easily told from the other taxa that grow in the Red Desert. Other Sagebrush Vegetation Several other types of sagebrush vegetation replace the Wyoming big sagebrush steppe on unusual habitats or at higher elevation around the margins of the desert.

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 209

basin big sagebrush Drive Highway 191 and then Sweetwater County Road 17 north from Rock Springs to the intersection by Fifteenmile Knoll Reservoir, continue northeast on the county road along Nitch Creek, and turn north on the small road through the Killpecker Dunes to Indian Gap. On the dunes you’ll see stands of sagebrush like no other in Wyoming, with shrubs so tall that they hide standing elk. This is basin big sagebrush (Artemisia tridentata ssp. tridentata), the tallest of the sagebrushes. Throughout the west, it grows on deep, relatively moist soils in the intermountain basins. Early white settlers recognized basin big sagebrush as an indicator of rich soil, and many stands around the west have been plowed for hay meadows or cropland. Basin big sagebrush grows throughout the western half of Wyoming as ribbons or small patches of tall shrubs along streams. The Killpecker Dunes stands, though, are different. Since the mid-1940s, when the Wyoming Game and Fish Department reintroduced elk to their old winter range in the western Red Desert, this tall basin big sagebrush shrubland has been recognized as crucial habitat for the animals.23 Elk are primarily grazers, not browsers (that is, they eat mainly grasses and forbs instead of shrubs), and so they depend less on sagebrush for food than do pronghorn, greater sage-grouse, pygmy rabbits, and even mule deer. Rather, they depend on the basin big sagebrush for cover. Studies of radio-collared elk have shown that they use tall sagebrush vegetation during the calving season from mid-May through June to hide their young, and in July and August for shade.24 This unusual shrub vegetation has sparse undergrowth, and wildlife managers have expressed concern that heavy grazing by the elk has reduced the amount of herbaceous plant cover below. Data collected during a survey of the tall sagebrush vegetation, though, showed no relationship between plant cover in the undergrowth and the number of elk droppings.25 But the amount of undergrowth cover is inversely related to both the height and the density of the shrub canopy, which suggests that the undergrowth is so sparse because the shrubs create an unfavorable environ-

ment. Sparse ground cover is a well-known feature of dense forests, and the basin big sagebrush vegetation on the dunes is, essentially, a low forest. The mechanism may be shading of the undergrowth plants by the shrub overstory, or the shrubs may be outcompeting the grasses and forbs for water. This remarkable sagebrush vegetation owes its existence to the sand of the Killpecker Dunes. The tall shrubs grow only where the wind-driven sand has buried the sedimentary bedrock. On outcrops of that finer-textured bedrock (marked by pieces of sandstone on the ground), the vegetation is short and the sagebrush is Wyoming big sagebrush. This dependence of the tall basin big sagebrush on the dune sand is clear also where the strong southwesterly winds have pushed sand over Steamboat Rim and into the heads of the canyons that run from there to the north. In those canyons, you can map the extent of the sand by noting where the tall sagebrush grows. From the county road between Steamboat Mountain and north Table Mountain too, you can see how far the sand has lapped up the side of Steamboat Mountain by the extent of the tall sagebrush. Sand is often thought of as a dry substrate, but in arid climates where evaporation from the soil surface to the atmosphere (rather than downward flow out of the soil) accounts for much of the loss of soil water, sand is a better water reservoir than is fine-textured soil. Water percolates deeper into sand, so more of it is stored below the reach of evaporation and remains available to plants. You can see this for yourself if you dig in the Killpecker Dunes: even in midsummer, you quickly go through the dry upper layers and reach the moist sand beneath. With its combination of concentrated shallow roots and sparser deep roots, a basin big sagebrush plant can extract enough water from the sand to support its large volume of living tissue. And it experiences dry-soil years, when it can barely grow, less often than do the plants on finer soils. The undergrowth of the tall basin big sagebrush stands contains plant species that are common in the Wyoming big sagebrush steppe, especially Indian ricegrass, needle-and-thread grass, and Douglas rabbitbrush. But two plant species that are rare in the

Sagebrush

209

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 210

sagebrush steppe—lemon scurfpea (Psoralidium lanceolatum) and rubber rabbitbrush (Ericameria nauseosa)—are so common in the tall sagebrush stands that they (and basin big sagebrush) represent this vegetation. Lemon scurfpea is a rhizomatous forb that contributes most of the cover to the sparse undergrowth in the tall sagebrush stands. Rubber rabbitbrush also is common and, in many places, contributes as much cover as does the sagebrush. Both scurfpea and rubber rabbitbrush sprout from underground parts and thus cope with at least shallow burial by sand. These sprouting plants might be expected on dunes. Basin big sagebrush, though, has no such adaptation for growing in active sand. So why is it the dominant plant on the dunes? Observations by the geologist Thomas Ahlbrandt provide the answer.26 He pointed out that basin big sagebrush dominates the vegetation on dunes that have been stabilized the longest. Dunes that have been active more recently are dominated by rubber rabbitbrush, and the most active dunes are unvegetated or support, at best, mats of lemon scurfpea. This relationship of vegetation to dune stability can be seen on many of the east-west – trending dune ridges in the area. The active tops of the ridges are bare, the upper sides of the dunes are green with rubber rabbitbrush, and the lower dune slopes and depressions between dunes are clothed in gray sagebrush. The tall sagebrush stands, then, might be considered the climax vegetation. A site is first occupied by lemon scurfpea and a few other rhizomatous forbs and grasses. Rubber rabbitbrush gradually moves in and forms a shrub overstory as the sand becomes stabilized. Once established, these earlier species remain on the site but are gradually overtopped by the basin big sagebrush. If the site is covered with sand, then the sequence starts again. A similar sequence of vegetation types develops when a site in the basin big sagebrush shrubland burns or is disturbed by construction of pipelines or well pads.27 No doubt the herbaceous layer and the rubber rabbitbrush layer develop more quickly on burns or small pipelines than on new sand dunes because many of the plants growing in the burned area or pipeline corridor be210

fore the disturbance sprout in the first year after the disturbance. On reclaimed well pads and wide pipeline corridors, though, the plants must start from seed and recovery is slower. It is unclear how much time is required for reestablishment of the basin big sagebrush overstory, but that probably happens rather slowly because the plants must start from seed after any sort of disturbance. When you sit in the shade of especially tall sagebrush and contemplate their twisted trunks, it’s tempting to think that you’re in the presence of venerable matriarchs whose lives have spanned much of the Red Desert’s history. And, relative to many plants, they may be old. Alan Beetle reported big sagebrush plants in southeastern Wyoming over 100 years old.28 But he also pointed out that the size of a sagebrush plant tells us more about the place where it grows than about its age. Large plants grow on sites where soil water usually is plentiful and they can add girth and height virtually every year. Plants on windswept slopes and flats, in contrast, frequently find so little soil water that they can hardly grow and are small even though they may be decades old. Surprisingly, Perryman and his co-workers found that plants of basin big sagebrush from stands around Wyoming were younger (an average age of 55 years) than mountain big sagebrush (81 years) or Wyoming big sagebrush (75 years).29 They noted, though, that the requirements of their sample design may have caused them to miss the oldest plants and thus skewed their results. Basin big sagebrush in the Red Desert also grows in smaller stands typical of the region, where deep alluvial soils in draws and along the few perennial streams provide large soil water reservoirs. The vegetation on these medium-textured soils is quite different from that on the Killpecker Dunes.30 Here, black greasewood and basin wildrye often grow in the overstory with the sagebrush, and mountain silver sagebrush (Artemisia cana ssp. viscidula) may be present as well. The undergrowth on alluvial soils usually contains greater amounts of rhizomatous grasses, and lemon scurfpea is absent. Along ephemeral streams, the basin big sagebrush typically fills the bottom of the draw. Along the larger, meandering streams, such as the Sweetwater River,

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 211

basin big sagebrush grows in patches on intermediate terraces. In the Jack Morrow Hills of the western Red Desert, you can see basin big sagebrush growing at the foot of east-facing escarpments, where snowdrifts form in the strong westerly winds. Bush Rim north of Steamboat Mountain is a good illustration of this phenomenon.31 The basin big sagebrush grows with mountain big sagebrush and basin wildrye in dense stands at least three feet tall. Utah snowberry (Symphoricarpos oreophilus ssp. utahensis) and antelope bitterbrush (Purshia tridentata), two sprouting shrubs, grow beneath. These belts of tall vegetation with palatable shrubs provide mule deer with cover and food in an otherwise open environment. silver sagebrush Like basin big sagebrush, silver sagebrush (Artemisia cana) can be found in two seemingly different types of habitats in the Red Desert: on alluvial soils in draws, and on sandy soils of stabilized dunes. The common feature of these habitats is a supply of soil water greater than that provided by precipitation to upland soils derived from sedimentary bedrock. Mountain silver sagebrush (Artemisia cana ssp. viscidula) is a common inhabitant of moist meadows and streamsides throughout the western United States. Such habitats are rare in the arid Red Desert, and much of the land there may be below this sagebrush’s elevation range. Consequently, plants of this mediumheight, deciduous, sprouting shrub with its narrow, dark green-gray leaves are restricted to the larger draws. If you travel Sweetwater County Road 21 between Rock Cabin Creek and Bush Rim, stop at the crossing of Parnell Creek. The shrubs in the bottom of the draw there are mountain silver sagebrush. In the right light, you can tell from the road that their leaves are darker than those of the Wyoming big sagebrush you’ve been seeing. The leaves are long and narrow and, although they may be lobed, lack the three terminal teeth of big sagebrush. The other subspecies of silver sagebrush, plains silver sage-

brush (Artemisia cana ssp. cana), is equally restricted in the Red Desert. This is the most common shrub in the Sand Hills north of Baggs, on the eastern edge of the Red Desert. The vegetation of the Sand Hills is a shrub steppe similar in herbaceous species composition to the Wyoming big sagebrush shrub steppe.32 The shrub layer is richer, though: silver sagebrush, Douglas and rubber rabbitbrush, antelope bitterbrush, and Utah snowberry are common. Basin big sagebrush is present but seems to grow only in patches. Why does silver sagebrush dominate in some of the sites with greater soil moisture and basin big sagebrush in others? The answer may be straightforward in respect to the alluvial soils in draws and along streams. Silver sagebrush can tolerate longer periods when soils are very wet and little oxygen is available to plant roots. Big sagebrush, in contrast, succumbs very quickly to flooding. Alluvial soils, particularly fine-textured soils, may be waterlogged for enough of the year to kill the intolerant big sagebrush but not the silver sagebrush. Certain soil features are diagnostic of the anaerobic environment in waterlogged soils, and soil pits dug into the alluvial soils would give the information necessary to test this hypothesis. Sandy upland soils, such as those in the Killpecker Dunes and the Sand Hills, are well drained, and waterlogging is highly unlikely there. Why, then, does silver sagebrush, not basin big sagebrush, dominate in the Sand Hills? Perhaps the answer is the opposite of water; perhaps it is fire. Much of the land in the Sand Hills is known to have burned at least once since 1968 (most recently in 1993). The common shrubs there (including silver sagebrush) sprout after fire, and they would be favored by fire over a nonsprouting shrub like basin big sagebrush. Despite the presence of sprouting shrubs, shrub cover is still less on the burned areas than on the areas unburned since 1968.33 Recovery of the shrubs is slow. mountain big sagebrush If you keep your eye on the vegetation as you travel in Wyoming,

Sagebrush

211

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 212

you’ll notice that the low-growing sagebrush plants of the basins are replaced in the surrounding foothills by taller, more robust sagebrush, often growing in dense stands. This is mountain big sagebrush (Artemisia tridentata ssp. vaseyana), the subspecies that grows with scattered limber pine on the open slopes and dominates the parks among aspen and fir groves. Throughout the west, mountain big sagebrush is known to favor deep, usually moist soils at higher elevations than Wyoming or basin big sagebrushes. The Red Desert is too low to be prime mountain big sagebrush habitat, but the plant grows there. In the Jack Morrow Hills, mountain big sagebrush dominates shrub patches in draws and (with basin big sagebrush and robust Wyoming big sagebrush) forms bands on east-facing slopes.34 The presence of Utah snowberry, antelope bitterbrush, serviceberry (Amelanchier alnifolia), and chokecherry (Prunus virginiana) brings to mind the mixedshrub vegetation of the mountain foothills, and indeed these shrub patches can be thought of as low-elevation outliers of that vegetation. Their small area on the landscape belies their importance, as they provide mule deer with food and cover in the otherwise open country. Here too is where one listens for the green-tailed towhee, a foothill bird. Patches of mountain big sagebrush also grow in the Rawlins Uplift. The subspecies is most common at the higher-elevation margins of the Red Desert. At the southern foot of the Ferris Mountains and in the Antelope Hills, stands grow in draws and on lee slopes in the matrix of the shorter sagebrushes.35 The largest areas of mountain big sagebrush lie along the western foothills of the Sierra Madre at the desert’s eastern limit and in the Pacific Creek–Little Sandy River area in the desert’s higherelevation northwestern corner.36 The Future of Sagebrush in the Red Desert Two sorts of concern are being voiced throughout the west about deterioration in the condition of sagebrush vegetation. One seems to apply to the sagebrush steppe that covers much of the Red 212

Desert, the other primarily to the denser stands around the margin of the desert. Wildlife biologists have long valued sagebrush vegetation as habitat for pronghorn, mule deer, and greater sage-grouse, and declining populations of the latter two have focused attention on deterioration of their habitats. More recently, the focus has broadened to include additional animals and plants that inhabit the sagebrush vegetation, and also phenomena less directly connected to habitat quality as traditionally understood. This broader focus is expressed by the phrase “decline of the sagebrush ecosystem,” where “decline” encompasses such phenomena as loss of native plants and animals, increased abundance of exotic species, conversion of native sagebrush vegetation to other cover types, increased or decreased (depending on the location) fire frequency, increased soil erosion, and fragmentation of large tracts of sagebrush vegetation by roads and industrial facilities. Historically, invasion by exotic plants, especially cheatgrass (Bromus tectorum), has been a primary cause of decline in the sagebrush ecosystem. A short-lived exotic that competes strongly against the native grasses for water, benefits from fire, and thrives under grazing, cheatgrass figured prominently in the ruin of much of the sagebrush vegetation of the Intermountain region, the dismal but fascinating story of which is told by West and Young.37 In the late nineteenth and early twentieth centuries, grazing by livestock over much of the sagebrush country converted the mix of native grasses and sagebrush to dense sagebrush with little undergrowth. Cheatgrass then spread into the depauperate shrublands, and the consequent increase in fires (dried cheatgrass is spectacularly flammable) converted the vegetation to a cheatgrass grassland or a shrub steppe of cheatgrass with rabbitbrush. Cheatgrass is common in the sagebrush vegetation on the Great Plains of northeastern Wyoming and in the Wind River and Bighorn basins, where it dominates in places. But fortunately that vegetation has not suffered the fate of the sagebrush steppe farther west.38 In the Red Desert and the other higher-elevation basins of Wyoming, cheatgrass is far less common, and resource man-

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 213

agers and biologists have comforted themselves with the idea that the climate in those basins is too cold for the weed to thrive. In a static world, this hypothesis, if true, would allow one to sleep well at night. But both the climate and the biota are dynamic. If temperatures rise, even just a few degrees, cheatgrass may find the high-elevation basins quite hospitable. And even if temperatures remain the same, cheatgrass probably will not. Populations of this species contain a great deal of genetic variation, and even if the populations in or near the Red Desert now lack a genotype acclimated to cold, such a genotype might appear in the area and thrive. Cheatgrass might never become a problem in the sagebrush vegetation of the Red Desert, but several other exotics already there are cause for worry: halogeton (Halogeton glomeratus), a succulent annual that concentrates oxalates in its tissue to poisonous levels; whitetop, or hoary cress (Cardaria species), a group of three perennial mustards that sprout from deep roots; and Russian thistle, or tumbleweed (Salsola species), famous for its ability to disperse seed as the wind rolls it across the landscape. All three grow throughout the Red Desert. Fortunately they are far more common on disturbed sites such as roads, pipeline corridors, well pads, and livestock wallows than in the relatively undisturbed sagebrush steppe or shrub vegetation. And (again, fortunately) none seem to have cheatgrass’s capacity to alter the structure and function of the ecosystem. Unfortunately, the disturbed habitats that they exploit will become far more common with the energy development planned for the Red Desert in the coming decades. Only enormous efforts at monitoring and control might keep these weeds from spreading in the new well pads, roads, and pipeline corridors and altering the sagebrush steppe and desert shrub vegetation throughout the country. A second concern, voiced primarily by wildlife biologists, is that certain sagebrush-dependent animals are losing the nutritious food on which they depend. This concern arose at least in part because mule deer populations have not recovered from harsh winters as they used to do. The explanation offered for this lack of recovery is that, with fire suppression, stands of old, slow-

growing shrubs with little nutritional value are no longer replaced by stands of young, fast growing, nutritious shrubs. There can be little doubt that sagebrush shrubs have grown slowly in the past decade, a period of drought in much of Wyoming. But the details of the relationship between shrub productivity and fire frequency are hard to establish. Fire frequency has been highly variable from place to place in sagebrush vegetation,39 and apparently there has been little or no research on fire frequency in the Red Desert. It seems reasonable to hypothesize that much of the Wyoming big sagebrush of the steppe grows in such dry areas that the vegetation produces little fuel and fire suppression likely has had little effect on fire frequency and, hence, on nutritional value of the shrubs. In contrast, fire suppression might have had a greater effect in the mountain big sagebrush stands growing at the higher elevations. The data with which to test such a hypothesis will be hard to collect. We can say with confidence that sagebrush will be part of the Red Desert well into the future. We can say with even more confidence that the sagebrush vegetation of the Red Desert will change. Vegetation is dynamic. Its composition and structure derive from the characteristics of the constituent plant populations, and those populations increase or decrease in response to change in climate, arrival of new plant species and pathogens, increase and decrease in numbers of herbivores, change in fire regimes, and creation of disturbed habitats. Those populations themselves evolve. Recently in North America, the influence of humans on the vegetation has grown enormously. The question about the sagebrush vegetation of the Red Desert is not, Will it change? The question, rather, is, Can we influence the change so that the sagebrush ecosystem retains viable populations of the plant and animal species in it and continues to provide the things that we value from it?

Sagebrush

213

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 214

Notes 1. “Flora” means the plant species that grow in a particular area, without regard for which are common or rare. “Vegetation,” in contrast, denotes not only the species present but also the relative amounts of each and the overall structure that they create. 2. Julius Nagy, “Wildlife Nutrition and the Sagebrush Ecosystem,” in The Sagebrush Ecosystem: A Symposium (Utah State University, Logan, 1979). 3. R. O. Bray, C. L. Wambolt, and R. G. Kelsey, “Influences of Sagebrush Terpenoids on Mule Deer Preference,” Journal of Chemical Ecology 17 (1991): 2053–2062. 4. T. L. Personius, Carl L. Wambolt, Jeffrey R. Stephens, and Rick G. Kelsey, “Crude Terpenoid Influence on Mule Deer Preference for Sagebrush,” Journal of Range Management 40 (1987): 84– 88. 5. Nagy, “Wildlife Nutrition.” 6. Dennis Knight, Mountains and Plains: The Ecology of Wyoming Landscapes (Yale University Press, New Haven, Conn., 1994). This book is indispensable for anyone interested in the natural history of Wyoming. 7. The abbreviation “ssp.” stands for “subspecies.” 8. Graminoids are the grasses and sedges. Forbs are the broad-leaved, herbaceous plants—the wildflowers. These two classes of plants (plus the shrubs) have long been recognized by North American rangeland ecologists because of their different adaptations to grazing and fire. 9. David Sturges, “Soil Water Withdrawal and Root Characteristics of Big Sagebrush,” American Midland Naturalist 98 (1977): 257–274. 10. B. L. Perryman, A. M. Maier, A. L. Hild, and R. A. Olson, “Demographic Characteristics of Three Artemisia tridentata Subspecies,” in Sagebrush Establishment on Mined Lands: Ecology and Research, Proceedings of the 2000 Billings Land Reclamation Symposium, Montana State University Reclamation Research Publication 00-01, edited by Gerald E. Schuman, Timothy C. Richmond, and Dennis R. Neuman (Bozeman, Mont., 2000). 11. K. B. Cawker, “Evidence of Climatic Control from Population Age Structure of Artemisia tridentata Nutt. in Southern British Columbia,” Journal of Biogeography 7 (1980): 237–248. 12. Wyoming Gap Analysis Project, “Land Cover for Wyoming” (University of Wyoming Geographic Information Science Center, Laramie, 1996), http://www.sdvc.uwyo.edu/24k/landcov.html. 13. F. W. Olsen and R. M. Hansen, “Food Relations of Wild FreeRoaming Horses to Livestock and Big Game, Red Desert, Wyoming,” Journal of Range Management 30 (1977): 17–20. C. Sundstrom, W. G. 214

Hepworth, and K. L. Diem, Abundance, Distribution, and Food Habits of the Pronghorn, Wyoming Game and Fish Commission Bulletin 12 (Cheyenne, 1973). E. Taylor, Pronghorn Carrying Capacity of Wyoming’s Red Desert, Wyoming Game and Fish Department Wildlife Technical Report 3 (Cheyenne, 1975). 14. See Gary P. Beauvais, “Vertebrate Wildlife of the Red Desert,” this volume. 15. Olsen and Hansen, “Food Relations.” 16. Knight, in Mountains and Plains, presents a very readable description of patterns in sagebrush vegetation around Wyoming. 17. Rick Miller, “Fire in the Sagebrush Biome: Past and Present and Plant Community Response,” in Improvement and Management of Sagebrush Communities in Wyoming (Wyoming Game and Fish Department, Cheyenne, 2003). 18. Sturges, “Soil Water Withdrawal.” 19. George P. Jones, “Inventory and Mapping of Plant Communities in the Ferris Mountains Wilderness Study Area, Carbon County, Wyoming” (unpublished report, Wyoming Natural Diversity Database, University of Wyoming, Laramie, 2004). George P. Jones, “Inventory and Mapping of Plant Communities in the Sweetwater Canyon Wilderness Study Area, Fremont County, Wyoming” (unpublished report, Wyoming Natural Diversity Database, University of Wyoming, Laramie, 2004). 20. D. M. Beale and A. D. Smith, “Forage Use, Water Consumption, and Productivity of Pronghorn Antelope in Western Utah,” Journal of Wildlife Management 34 (1970): 570 –582. 21. Personius et al., “Crude Terpenoid Influence.” 22. B. Behan and B. L. Welch, “Black Sagebrush: Mule Deer Winter Preference and Monoterpenoid Content,” Journal of Range Management 38 (1985): 278 –280. 23. Dewey Henderson, “The Sands,” Wyoming Wildlife 19, no. 7 (1955): 4– 6. For a more recent discussion of the importance of the tall sagebrush to the unusual desert elk herd, see Dave Lockman, “Mama Sage’s Elk Herd,” Wyoming Wildlife 17, no. 2 (1978): 25–27. 24. Jacob H. Powell, Distribution, Habitat Use Patterns, and Elk Response to Human Disturbance in the Jack Morrow Hills, Wyoming (M.S. thesis, University of Wyoming, Laramie, 2003); Hall Sawyer and Ryan Nielson, Seasonal Distribution and Habitat Use Patterns of Elk in the Jack Morrow Hills Planning Area, Wyoming (Western Ecosystem Technology, Cheyenne, 2005), http://www.west-inc.com/big_game_reports.php. 25. George P. Jones, “Survey of Tall Sagebrush Vegetation on Stabilized Sands in the Jack Morrow Hills Coordinated Management Area, BLM Rock Springs Field Office, Wyoming” (unpublished report,

George P. Jones

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 215

Wyoming Natural Diversity Database, University of Wyoming, Laramie, 2005). 26. Thomas Ahlbrandt, “Sand Dunes, Geomorphology and Geology, Killpecker Creek Area, Northern Sweetwater County, Wyoming” (Ph.D. diss., University of Wyoming, Laramie, 1978). 27. Jones, “Tall Sagebrush Vegetation.” 28. Alan A. Beetle, A Study of Sagebrush: The Section Tridentatae of Artemisia, University of Wyoming Agricultural Experiment Station Bulletin 368 (Laramie, 1960). 29. Perryman et al., “Demographic Characteristics.” 30. George P. Jones and Walter Fertig, “Plant Associations and Plant Species of Special Concern in the Jack Morrow Hills Ecosystem” (unpublished report, Wyoming Natural Diversity Database, University of Wyoming, Laramie, 1995). 31. Ibid. 32. George P. Jones, “Vegetation on Sand Substrates in the BLM Rawlins and Rock Springs Field Offices, Wyoming” (unpublished report, Wyoming Natural Diversity Database, University of Wyoming, Laramie, 2005). 33. Ibid. 34. Jones and Fertig, “Plant Associations.” 35. Jones, “Ferris Mountains.” Jones, “Sweetwater Canyon.” 36. Wyoming Gap Analysis Project, “Land Cover for Wyoming.” 37. Neil E. West and James A. Young, “Intermountain Valleys and Lower Mountain Slopes,” Chap. 7 in Michael G. Barbour and William Dwight Billings, eds., North American Terrestrial Vegetation (Cambridge University Press, New York, 2000). 38. The explanation for the greater resilience may lie in the difference in grazing histories of the regions. The plants of the Wyoming basins and the Great Plains evolved under pressure from countless grazing animals, while the plants of the Intermountain region faced little grazing pressure until the very recent arrival of domestic livestock. This hypothesis is discussed by R. N. Mack and J. N. Thompson, “Evolution in Steppe with Few Large, Hooved Animals,” American Naturalist 119 (1982): 757– 773. 39. Estimates of intervals between fires in sagebrush vegetation before settlement by Europeans range from twelve to twenty-five years in moist, productive sites to thirty-five to one hundred years in dry sites. Miller, “Fire in the Sagebrush Biome.”

Sagebrush

215

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 216

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 217

10. Bright Green Hues Are Rare Plant Diversity and Conservation in Wyoming’s Red Desert Walter and Laura Fertig

Nothing but wild sage and the greasewood shrub were to be seen all over the country. . . . Since leaving the Platte we had again passed through a dreary, hilly country, . . . meeting with nothing in the least interesting to the traveler, being destitute of vegetation, except the Artemisia, which was seen from the highest hill to the lowest valley. . . . A more barren region cannot be found between here and the Columbia River. —major osborne cross, u.s. army, crossing the red desert in 1849 There is little in view from the car windows to invite closer inspection or more personal contact. From the Platte River to the Green River, a distance of 150 miles, there is not within sight of the railroad any vegetation larger than sage-brush and grease-wood, and through much of the distance these also seemed dwarf and scattering. Bright green hues are extremely rare. —aven nelson, university of wyoming botany professor, crossing the red desert by train in 1897

travel-weary forty-niner might be forgiven for overlooking the flora’s subtle charms, but one knows the Red Desert had an image problem when Aven Nelson, the state’s preeminent botanist, was unmoved. Like many desert lands, the Red Desert—a general term for the inwardly draining Great Divide basin and adjacent mountains and lowlands of south-central Wyoming—is a decidedly acquired taste.1 Most travelers, whether following the California-Oregon-Mormon Trail, Union Pacific, or Interstate 80, perceive an unremarkable terrain with low hills and broad valleys of drab brownish gray mudstone and sandstone. The thin mantle of vegetation, consisting largely of short, grayish, often spiny shrubs rather than green trees or flower-dotted meadows, does little to hide the sere landscape. To the casual observer in a hurry to reach someplace else the desert is dull, uninviting, and uncomfortable—hot in summer, cold in winter, and dry, windy, and dusty all the time. Little wonder then that the Red Desert has always been over-

A

shadowed by Wyoming’s more conventionally inspiring natural features such as Yellowstone, the Wind River Mountains, and Devils Tower. These were the lands deemed worthy of protection as national parks, monuments, and wilderness areas. The Red Desert, like so much of Wyoming’s basin country, was relegated to the railroad and the General Land Office (later the Bureau of Land Management, or BLM) and became the domain of a small number of shepherds, cattlemen, prospectors, desert rats, and others drawn to the outer margins of civilization. Despite a lack of formal protection, the Red Desert functioned as a de facto sanctuary for wildlife and native plants well into the 1990s. This era of benign neglect came to a screeching halt with the discovery of two increasingly rare and tremendously valuable commodities: untapped energy resources and vast open space for recreation. Conservationists who had largely ignored the Red Desert for decades now found themselves celebrating its virtues and scrambling for its protection. 217

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 218

What are we at risk of losing with the commodification of the Red Desert? Usually wildlife comes to mind before anything else. But the first order in saving wild animals is preserving vegetation—the source of their food, shelter, and cover. Native plants provide other critical ecological and economic services, such as reducing soil erosion, retaining moisture, storing carbon, and feeding livestock and humans. Plants also constitute a large part of the total biodiversity of the Red Desert (second only to invertebrates), but a high proportion of the flora is not protected in Wyoming’s existing network of national parks and wilderness areas. In short, effective conservation of the Red Desert needs to start with native plants. Plant Diversity in the Red Desert Aven Nelson conducted the first formal botanical survey of the Red Desert, documenting 173 species of vascular plants in the vicinity of the Union Pacific Railroad in 1897.2 Since then, over 900 additional species and varieties have been recorded, bringing the known flora to 1,074 taxa.3 This represents about 38 percent of the total flora of Wyoming.4 Although surpassed by Yellowstone National Park and the Absaroka and Big Horn ranges, the Red Desert has a larger flora than most of Wyoming’s other mountain ranges, grasslands, and desert basins. The diversity of plant species in the Red Desert far exceeds that of vertebrates, of which only 389 resident fish, amphibian, reptile, bird, and mammal species have been reported.5 The high species diversity of the Red Desert is a consequence of its varied topography, geology, and geographic location. More than thirty vegetation types have been documented, including coniferous forests, sagebrush grassland, greasewood and saltbush shrublands, sand dunes, and alkaline wetlands.6 The Red Desert’s proximity to three major biogeographic provinces results in additional diversity.7 Nearly one-half of the flora consists of desert species found widely across the Great Basin province, which dominates the basin and range country between the Sierra Nevadas 218

and Rocky Mountains. Approximately 20 percent of the plant taxa are perennial grasses and forbs found commonly in the Great Plains of central North America. Components of the Rocky Mountain cordilleran province (30 percent of the flora) occupy the higher elevations of Crooks, Green, and Ferris mountains along the northern rim of the Great Divide basin. About 85 percent of the species of the Red Desert occur widely across western North America, often in two or more biogeographic provinces. The rest of the flora is made up of 13 species restricted to Wyoming (state endemics), 74 species found in a limited geographic area centered on Wyoming and one or two adjacent states (regional endemics), and 42 species at the margin of their continental range in Wyoming (peripherals).8 Many of these plants are of conservation interest due to their small population size, high habitat specificity, or limited range. The majority of state and regional endemics of the Red Desert occur in badlands, sagebrush grassland, or desert saltbush habitats on poorly developed soils or harsh geologic substrates (especially limestones and shales). Botanist Robert Dorn has speculated that many of Wyoming’s desert endemics evolved from species of cooler and moister climates that were isolated and forced to adapt as the Wyoming basins became more aridic at the close of the Pleistocene.9 Species diversity in the Red Desert has been augmented in the last two centuries by the arrival of 100 nonnative plant taxa (representing just over 9 percent of the entire flora). A few of these aliens were introduced intentionally as ornamentals, crops, or livestock forage, including Russian olive (Elaeagnus angustifolia), red and white clover (Trifolium pratense and T. repens), alfalfa (Medicago sativa), smooth brome (Bromus inermis), crested wheatgrass (Agropyron cristatum), Kentucky bluegrass (Poa pratensis), and several other range grasses. Most of the other exotics arrived inadvertently from Europe or Asia through contaminated seed, hay, or packing material. Only about half of the introduced species have become widely established in Wyoming, and most of these are limited to frequently disturbed habitats on roadsides, irriga-

Walter and Laura Fertig

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 219

tion ditches, or agricultural fields. A small number, however, have become significant pests and now dominate riparian areas, sagebrush grasslands, or even clay barrens. Among the most deleterious of the new arrivals are Russian thistle (Salsola tragus), cheatgrass (Bromus tectorum), and halogeton (Halogeton glomeratus), which tend to displace edible natives, increase the likelihood of fire, or (in the case of halogeton) poison livestock. Major Habitats and Plant Adaptations of the Red Desert Sagebrush grassland is the most widespread vegetation type in the Red Desert, covering nearly three-quarters of the area. Although often considered floristically depauperate, nearly 60 percent of the desert flora has been documented in sagebrush habitats.10 The aptly named Wyoming big sagebrush (Artemisia tridentata ssp. wyomingensis) is the most common of the desert sagebrushes and thrives on fine-textured soils of the broad basins and low ridges found across the Red Desert. In other environments it is replaced by different taxa: basin big sagebrush (A. tridentata ssp. tridentata) on deeper, sandier soils; mountain big sagebrush (A. tridentata ssp. vaseyana) at higher elevations; silver sagebrush (A. cana) in riparian areas; black sagebrush (A. nova) on shallow, carbonate-rich soils; and alkali sagebrush (A. arbuscula ssp. longiloba) on granitic-derived clays.11 Sagebrush possesses several physical and physiologic adaptations that allow it to thrive in droughty environments. It has a shallow root system to take advantage of surface moisture early in the growing season and deeper roots that capture groundwater later in the summer. Many Artemisia species produce large, ephemeral leaves in the spring to maximize photosynthesis when moisture is plentiful. These leaves are shed in summer when water becomes limited, and the plant relies instead on smaller, evergreen leaves that are more drought-hardy and can continue to function through the winter. The characteristic bluish green color of sagebrush comes from closely matted hairs that help shield the leaf surface from the sun and reduce evaporative water loss. The

narrow leaves reduce the amount of surface area to minimize evaporation and are covered by pores (stomates in botanical jargon) that can be closed rapidly to retain moisture. At least 50 species of perennial grasses occur in sagebrushdominated habitats of the Red Desert. About 90 percent of these are “cool season,” or C3, grasses that flower and complete most of their new growth in spring and early summer when soil moisture is readily available.12 Cool-season grasses include some of the most important forage species for wildlife and livestock, such as Indian ricegrass (Achnatherum hymenoides),13 needle-and-thread (Hesperostipa comata), basin wildrye (Elymus cinereus), western wheatgrass (E. smithii), and various species of bluegrass (Poa). These are also the grass species that are most vulnerable to late spring or summer grazing and that may be less abundant today than they were prior to settlement.14 Warm-season, or C4, grasses are more efficient at using water for photosynthesis than coolseason species and are able to reproduce and remain metabolically active during the hottest and driest months of midsummer (coinciding with periods of low livestock use).15 Despite these advantages, warm-season grasses are rarely encountered in the Red Desert, suggesting that moisture is less of a limiting factor here than in hotter deserts of the southwest and Mexico, where C4 species are often ecologically dominant.16 Although often not abundant, perennial forbs (perhaps better known in common parlance as “wildflowers”) greatly exceed grasses in terms of species diversity in sagebrush habitats. Over 300 species of perennial forbs (or about one-half of the total flora of the Red Desert) have been documented in sagebrush grasslands. Whereas perennial grasses rely on rapid growth and efficient water use for survival, many perennial forbs grow slowly and devote significant food reserves to relatively long-lived leaves, woody stems, and thick underground roots, bulbs, or corms. With such an important investment to protect, perennial forbs frequently develop sharp spines or chemical defenses to render their foliage unpalatable (or even poisonous) to herbivores. Forbs tend to have larger leaves than grasses, a good adaptation for photo-

Bright Green Hues Are Rare

219

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 220

synthesis when water is not scarce in early spring, but a liability later in the summer. Thus many desert forbs lose their leaves or become dormant later in the year (in extreme cases even riding out the hottest months below ground). Most perennial desert forbs synchronize their flowering to coincide with periods of adequate moisture and warm (but not excessively hot) temperatures. For an all too brief period in late May and early June, the normally drab sagebrush grasslands of the Red Desert can erupt in a tumult of yellow, white, orange, and blue flowers, each advertising its availability to passing bees, butterflies, and hummingbirds. In basin areas with saline and clayey soils, sagebrush grassland gradually gives way to desert shrublands dominated by various species of saltbush, greasewood, and other shrubby members of the goosefoot family (Chenopodiaceae). Saltbush and greasewood communities support less than one-third the number of plant species found in sagebrush grasslands. Different chenopodiaceous species dominate depending on the aridity and salinity of a site. Arid lowlands with fine-textured, saline soils derived from shale or alluvium are frequently dominated by Gardner saltbush (Atriplex gardneri). This low-growing subshrub is especially valuable as a winter forage species for domestic sheep and is occasionally used in reclamation. Gardner saltbush is replaced by shadscale (Atriplex confertifolia) or winterfat (Krascheninnikovia lanata) on nonsaline, sandier soils.17 Shadscale is a C4 species and so is able to use water more efficiently during the summer months than winterfat, a C3 species. Greasewood (Sarcobatus vermiculatus) becomes dominant in playas, depressions, and the margins of ponds and lakes where the water table is shallow and soils are highly saline, sodic, or alkaline.18 The current distribution of greasewood vegetation in the Red Desert is strongly correlated with the ancient boundaries of Quaternary Lake Wamsutter, which deposited fine, silty, saline soils. Greasewood is able to survive in salty soils, where other plants would lose precious water via osmosis, because of its ability to tolerate high solute concentrations in its cells and to pump excess salts out of its foliage through specialized salt glands. 220

Extensive areas of the Red Desert consist of sparsely vegetated badlands and wind-blasted ridgecrests. Fewer than 10 percent of desert plant species are capable of surviving under these adverse conditions. Those that can survive are often annual forbs and grasses that germinate in winter or early spring, when snowmelt or spring precipitation is greatest. Annuals flower and set seed in one season, usually completing their life cycle just as the drought conditions of summer arrive. Stony ridgecrests are mostly inhabited by low-growing, densely matted perennial forbs commonly called cushion plants due to their resemblance to a vegetable pillow. Constant high winds and blowing grit shred any leaves or stems that protrude more than a few centimeters above the ground. Another advantage of the matted growth form is to minimize the effects of extreme temperatures and drying winds. Flowers, too, are usually borne near the ground, giving cushion plant communities the appearance of a floral carpet when in full bloom. Wind is also a prominent feature of the Killpecker Dunes, one of the largest sand dune complexes in North America. Shifting sand is an especially harsh environment, as it can quickly bury and suffocate plants or expose delicate roots and stems to desiccation. Most plants that colonize dunes produce extensive networks of semiwoody belowground stems (rhizomes) that help the plant maintain its position in a shifting dune and exploit a larger area for soil moisture. These rhizomes send up new shoots to replace those that are buried or damaged by the wind. Sand is also a relatively infertile substrate, offering little in the way of nutrients for plant growth. Many dune species harbor nitrogen-fixing bacteria in specialized root structures that gradually improve the nutrient-deficient soils. Cover and species diversity of active dunes is often low, consisting mostly of Indian ricegrass, needle-andthread, western wheatgrass, rusty lupine (Lupinus pusillus), and lemon scurfpea (Psoralidium lanceolatum). Eventually some dunes stabilize and can be colonized by shrubs such as big sagebrush, shadscale, rabbitbrush (Ericameria or Chrysothamnus species), and spiny hopsage (Grayia spinosa). Stable dunes have higher vegeta-

Walter and Laura Fertig

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 221

tive cover and species diversity. Dune stability can be reversed, however, through drought, fire, vehicle damage, and heavy grazing that reduce cover and expose fresh sand to the sculpting forces of the wind. Overall, nearly 250 plant species have been recorded in the Killpecker Dunes. Perhaps the most incongruous of all desert habitats are wetlands. Although limited in area, the ponds, streams, and springs of the Red Desert provide habitat for 315 vascular plant taxa (onethird of the flora) and nearly all aquatic and terrestrial vertebrate species. Most riparian plants are not especially adapted to arid desert environments and die out during extended drought or if water sources are diverted. The traditional riparian vegetation of willows, cottonwoods, birches, and dense grasses is mostly restricted to the Sweetwater River and its main tributaries. Within the heart of the Red Desert, most wetlands are found only in lowlying depressions where water drains from surrounding uplands through coarse sand or along an impervious clay layer. Although short-lived ponds may form in greasewood- or saltbushdominated playa bottoms after heavy rains, true riparian vegetation develops only where water sources are consistently reliable. Small ponds occur frequently in the Killpecker Dunes, where meltwater from buried snow or infrequent showers accumulates in interdunal sand depressions. Such communities are usually composed of Baltic rush ( Juncus balticus), foxtail barley (Hordeum jubatum), three-square bulrush (Schoenoplectus pungens),19 Nebraska sedge (Carex nebrascensis), and other graminoids.20 More often, wetlands in the Red Desert are on alkaline clays. Different vegetation types form concentric rings around a water source according to a gradient of increasing dryness and decreasing salinity. The innermost ring consists of emergent marshlands of sedges and bulrushes on flooded or saturated soils. Rocky Mountain glasswort (Salicornia rubra) and other annual or herbaceous chenopodiaceous forbs occur on slippery clay barrens just above the water’s edge. Many of these species are halophytes capable of pumping out excess salts from their semi-succulent stem tissues, just like greasewood. Meadows of inland saltgrass (Dis-

tichlis stricta) and alkali sacaton (Sporobolus airoides) form the next zone, where soils remain moist but are slightly less alkaline. Greasewood shrublands typically form the outer ring of the wetland zone and can even grow on terraces if their long taproots reach the water table. Shrubby desert vegetation gives way to tree-dominated woodlands in the foothills of mountain ranges forming the northern boundary of the Red Desert. Temperatures and precipitation remain desertlike in this transition zone, but soils change abruptly from deep, fine-textured clay or sand in the basins to shallow, coarse, rocky soils or exposed bedrock on the lower mountain slopes. Because of their longer taproots, trees are better suited than shrubs or grasses to exploit water held in rocky soils or bedrock cracks. Limber pine (Pinus flexilis) and Utah juniper (Juniperus osteosperma) are the dominant tree species in the foothills zone. Both of these evergreen conifers are able to photosynthesize year-round, whenever temperatures are high enough, aided by resins within the leaf tissue that act as an antifreeze in winter. Their needlelike or scalelike leaves are protected from desiccation through reduction of surface area, deeply sunken stomates, and waxy coatings. Though expensive to produce, these leaves are durable and long-lived, and ultimately cheaper to manufacture than the leaves of deciduous trees that are produced and discarded every year. Limber pine and juniper typically form open woodlands, with an understory of shrubs, grasses, and perennial forbs commonly found in sagebrush grasslands. Diversity of these forests is relatively high, with just over 400 taxa documented. At higher elevations temperatures decrease, more water is available, and foothills woodlands are replaced by denser stands of limber pine, lodgepole pine (Pinus contorta), Douglas fir (Pseudotsuga menziesii), and Rocky Mountain juniper ( Juniperus scopulorum). This mixed conifer community closely resembles dry phases of lodgepole pine forests of the Wind River Range. Due to their isolation, aridity, and small area, the mixed conifer stands of the Crooks, Green, and Ferris mountains are relatively depauperate, with only 150 species recorded.

Bright Green Hues Are Rare

221

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 222

Small stands of aspen (Populus tremuloides) occur sporadically in the mountains and on some of the lower mesas and buttes of the Red Desert, including Steamboat Mountain and Bush and Joe Hay rims. As a broad-leaved deciduous tree, aspen requires more water than other woody forest or desert shrub species and is limited to springs or pockets where snow accumulates. Aspen groves are mini-oases of cool shade, loamy soils, and rich beds of wildflowers, grasses, and shrubs, few of which occur elsewhere in the desert. Not surprisingly, aspen stands tend to attract wildlife, livestock, and humans. Aspens reproduce primarily by root suckers, with individual trees living for about one hundred years. Heavy browsing or lack of fire can suppress the survival of new stems, eliminating young plants that would otherwise replace their aging parents. Although only about 140 plant taxa are found in aspen stands of the Red Desert, the loss of this habitat would result in a decline of nearly 10 percent of the area’s floral diversity. Changes in the Flora At first glance, the Red Desert’s vast expanses of sagebrush and saltbush seem to have changed little over time. Robert Dorn analyzed reliable written accounts of emigrants and explorers in the Red Desert from 1805 to 1872 and concluded that such desert mainstays as sagebrush and cactus were as common then as today.21 But other changes have occurred, both obvious and subtle. Many of the game species and predators commonly reported by early travelers (such as bison, bighorn sheep, beaver, grizzlies, wolverines, and sharp-tailed grouse) are now extirpated or extremely rare.22 Roads and pipelines now crisscross the desert along with oil and gas developments and scattered outposts of urbanization. At least 100 nonnative plant species have become established, and some (most notably cheatgrass, Russian thistle, and halogeton) have displaced native plants or significantly increased litter and fire hazards. But perhaps the most pervasive change has been the introduction of nonnative grazing animals— sheep, cattle, and wild horses. 222

In the late 1890s the Red Desert was primarily used as winter pasture for large herds of sheep. Between 300,000 to 800,000 stock were wintered in central Wyoming, many coming from as far away as Idaho, Utah, and Colorado.23 Sheep grazing continued to predominate into the 1920s. In 1926 Aven Nelson returned to the Red Desert to conduct a follow-up to his 1897 floristic survey for the U.S. Department of Agriculture. Specifically, Nelson wanted to assess whether range conditions had deteriorated on public lands of central Wyoming, as had been suggested by W. C. Barnes, chief of grazing for the U.S. Forest Service.24 Although he did no formal, quantitative analysis, Nelson concluded that the quality and abundance of edible shrubs and forage grasses had not materially degraded in the intervening thirty years and that any short-term changes were a response to climate and drought rather than grazing pressure.25 In a letter on behalf of the wool growers, Nelson reiterated a point he originally made in 1898: “The effectiveness of this browsing is, of course, dependent upon the region being used as a winter pasture only, giving time for growth and recovery each summer.” The face of grazing in the Red Desert changed in the 1930s. Passage of the Taylor Grazing Act in 1934 resulted in creation of individual permitted allotments in areas that had once been common pasture to all. Economics resulted in a shift from sheep grazing to cattle grazing and increased use of summer range. The ascendancy of automobiles resulted in many horses being abandoned to a feral existence, leading to increased, year-round pressure on rangelands from wild herds. Relatively few quantitative studies have been undertaken on the effects of grazing on the vegetation of the Red Desert. A twenty-year study comparing ungrazed relict areas on Pilot and Black buttes with adjacent grazed lands corroborated some of Nelson’s general observations about yearly fluctuations associated with climate, but also documented shifts in density and abundance of several important range species.26 The diversity of perennial forbs, total ground cover, and cover of big sagebrush were higher in grazed plots than in ungrazed sites, while the cover of

Walter and Laura Fertig

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 223

winterfat and perennial grasses was reduced. Another study of Red Desert summer and winter pastures grazed by horses and cattle found that both species preferentially used perennial coolseason grasses and winterfat over forbs, annuals, and other lesspalatable shrubs.27 Studies elsewhere in the west have found that the cover and abundance of unpalatable woody species (such as sagebrush and rabbitbrush) and annual weedy forbs have tended to increase with grazing, while edible cool-season grasses have declined.28 Red Desert Conservation Bigger changes may be in store for the vegetation of the Red Desert as industrial mineral and recreation development increase. The largely laissez-faire management approach of the past is no longer sufficient for the persistence of desirable native wildlife and plant species in the face of new and competing uses. Two changes in management are necessary to ensure the long-term survival of the native biota of the Red Desert. First, public land agencies need to give wildlife and the native flora equal footing with other natural resources in developing and implementing plans for multiple-use management. The view that multiple use means “all uses everywhere” needs to be amended to ensure that no single use overrides other values and that, at a minimum, the health and functioning of native plant communities are not compromised for short-term gain. Second, additional areas of significant plant and animal habitat need to be formally recognized and conserved. At present, less than 1 percent of the Red Desert is managed specifically and permanently for wildlife and habitat protection.29 A small number of Areas of Critical Environmental Concern have been created by the BLM to recognize places of significance to wildlife and rare plants, but these often lack specific management prescriptions and permanent long-term protection.30 Statewide, Wyoming’s existing network of protected areas is largely biased toward alpine and montane habitats of the greater Yellowstone region, with the

floras of desert basins, low-elevation wetlands, and grasslands poorly represented.31 Nearly all of the major upland vegetation types of the Red Desert have less than 1 percent of their total area under protection, and several (including active sand dunes, black sagebrush, basin big sagebrush, and unvegetated playas) are not protected at all.32 Nearly one-quarter of the native plant species in the Red Desert are unprotected or have less than 5 percent of their total range in a protected area. Giving native plants and vegetation equal footing with other, more traditional desert resources may seem a radical concept to some, but it is long overdue. For too many years we have taken the role of plants in ecosystems for granted. Without plants there is no habitat for game species, songbirds, and other wildlife we enjoy. There is no primary productivity, nutrient cycling, carbon storage, or other ecosystem services. There is no food for livestock, no new food crops for people,33 no new medicines, no colorful wildflowers to admire. But the most radical notion of all may be that plants and vegetation warrant preservation simply for their own sake—not for us, or wildlife, or livestock—but simply because they have intrinsic value. This perspective is perhaps a new one in the west. But it is a perspective to keep in mind as we, with our vast powers to alter the landscape and its constituent ecosystems, move further into this Red Desert country.

Bright Green Hues Are Rare

223

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 224

Notes 1. See Map 1, this volume. 2. Nelson, Red Desert. 3. “Taxon” (plural: “taxa”) is a general term for any described taxonomic entity, whether a full species, a subspecies, or a variety. The number of plant taxa in the Red Desert was derived from L. Welp, A floristic survey of the Great Divide basin, Green Mountains, and Upper Sweetwater Plateau in southwest Wyoming (master’s thesis, Department of Botany, University of Wyoming, Laramie, 1997); B. Ward, A floristic inventory of south-central Wyoming (master’s thesis, Department of Botany, University of Wyoming, Laramie, 1998); and the digital plant distribution atlas of the Rocky Mountain Herbarium, http://www.rmh .uwyo.edu/. 4. The Wyoming flora contains 2,800 documented taxa based on R. D. Dorn, Vascular Plants of Wyoming, 3rd ed. (Mountain West Publishing, Cheyenne, 2001). 5. The number of vertebrate taxa for the Red Desert is derived from E. H. Merrill, T. W. Kohley, M. E. Herdendorf, W. A. Reiners, K. L. Driese, R. W. Marrs, and S. H. Anderson, The Wyoming Gap Analysis Project Final Report (University of Wyoming, Laramie, 1996); and G. T. Baxter and M. D. Stone, Fishes of Wyoming (Wyoming Game and Fish Department, Cheyenne, 1995). 6. Important vegetation types of the Red Desert are summarized in D. H. Knight, R. J. Hill, and A. T. Harrison, Potential Natural Landmarks in the Wyoming Basin (University of Wyoming, Laramie, 1976); Merrill et al., Wyoming Gap Analysis Project; Welp, Floristic survey; G. Jones and W. Fertig, Plant Associations and Plant Species of Special Concern in the Jack Morrow Hills Ecosystem (Wyoming Natural Diversity Database, Laramie, 1996); and A. F. Vass and R. Lang, Vegetation Composition, Density, Grazing Capacity, and Grazing Land Values in the Red Desert Area, University of Wyoming Agricultural Experiment Station Bulletin 229. 7. Provinces are large geographic regions characterized by similar plant and animal species but are distinct from adjacent geographic areas. We follow the province characterization of R. G. Bailey, Descriptions of Ecoregions of the United States, Miscellaneous Publication 1391, U.S. Department of Agriculture (Washington, D.C., 1995). 8. Data on statewide endemism and plant geography derived from W. Fertig, R. Thurston, and W. A. Reiners, Modeling the Potential Distribution of Plant Species in Wyoming: A Tool for Gap Analysis (Department of Botany, University of Wyoming, Laramie, 2002).

224

9. R. D. Dorn and R. W. Lichvar, Vascular plants, in Clark and Dorn, eds., Rare and Endangered Vascular Plants and Vertebrates of Wyoming (Cheyenne, 1981). 10. Data on species diversity in each major habitat type of the Red Desert is derived from Welp, Floristic survey. 11. See G. P. Jones, Sagebrush, this volume, for a more thorough treatment of sagebrush ecology in the Red Desert, and D. H. Knight, Mountains and Plains: The Ecology of Wyoming Landscapes (Yale University Press, New Haven, Conn., 1994), for a discussion of sagebrush and other plant communities across the state. 12. C3 plants are so-named because the first stable product produced during photosynthesis is a sugar containing three carbon atoms. 13. Nomenclature for grasses and other plants in this chapter follows Dorn, Vascular Plants of Wyoming. To the chagrin of many, Dorn adopted new generic names for several familiar grasses, including Oryzopsis hymenoides and Stipa comata. Changes in nomenclature may appear arbitrary but actually reflect advances in taxonomic research and our understanding of species’ relationships and evolution. 14. Knight, Mountains and Plains; and R. Marquiss and R. Lang, Vegetational composition and ground cover of two natural relict areas and their associated grazed areas in the Red Desert of Wyoming, Journal of Range Management 12 (1959): 104–109. 15. C4 plants produce a four-carbon sugar as the first stable product during photosynthesis. C4 species are able to sequester more carbon in their cells and keep the concentration of CO2 gas low in the air spaces inside their leaves, allowing CO2 to be more readily taken up from the atmosphere. This means that the leaf stomates do not have to stay open for gas exchange during hot weather, reducing evaporative water loss and thus giving these species a competitive advantage in desert environments where water is limited. 16. According to Welp, Floristic survey, and Nelson, Red Desert, the main C4 grasses of Wyoming (Bouteloua gracilis, Hilaria jamesii, Muhlenbergia spp., Distichlis stricta, Spartina gracilis, Sporobolus airoides, and Sporobolus cryptandrus) are rarely found in the Red Desert except in wet meadow sites. 17. Winterfat has changed genus names several times over the years, going from Eurotia to Ceratoides and now Krascheninnikovia, which has the distinction of being the longest genus name in the flora of Wyoming. 18. Technically, greasewood is no longer considered a member of the Chenopodiaceae but instead belongs in its own family, the Sarcobataceae, based on differences in chloroplast DNA and sieve element anatomy.

Walter and Laura Fertig

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 225

19. Formerly placed in the genus Scirpus. 20. “Graminoid” is a catchall term for members of the grass, sedge, and rush families with linear, grasslike leaves and small, nondescript, wind-pollinated flowers. 21. Dorn, Wyoming Landscape. 22. Ibid. See also G. P. Beauvais, Vertebrate wildlife of the Red Desert, this volume. 23. Nelson, Red Desert. 24. Nelson’s visit was at the request of the Wyoming Wool Growers Association, which questioned Barnes’s critique of the management practices of the Grazing Service (a precursor of the BLM). Like so many other issues related to grazing, the Barnes debate was as much about politics as land management. Some believed that Barnes was more interested in shifting the responsibilities of the Grazing Service from the Department of Interior to his agency (the U.S. Forest Service) in the Department of Agriculture. 25. A. Nelson, The forage of the red desert (letter in the archives of the Rocky Mountain Herbarium, University of Wyoming, Laramie). 26. Marquiss and Lang, Vegetational composition. 27. L. J. Krysl, M. E. Hubbert, B. F. Sourell, G. E. Plumb, and T. K. Jewett, Horses and cattle grazing in the Wyoming Red Desert, I: Food habits and dietary overlap, Journal of Range Management 37 (1984): 72 –76. 28. Knight, Mountains and Plains; and R. F. Miller, T. Svejcar, and N. West, Implications of livestock grazing in the intermountain sagebrush region: Plant composition, in M. Vavra, W. A. Laycock, and R. D. Pieper, eds. Ecological Implications of Livestock Herbivory in the West (Society for Range Management, Denver, 1993). 29. Merrill et al., Wyoming Gap Analysis Project. The one area with permanent protection is the Sweetwater River Preserve, managed by the Wyoming Nature Conservancy. 30. W. Fertig, L. Welp, and S. Markow, The Status of Rare Plants in Southwest Wyoming (Wyoming Natural Diversity Database, Laramie, 1998). 31. W. Fertig and R. Thurston, Gap analysis of the flora of Wyoming, Gap Analysis Bulletin 10 (2001): 3– 6. 32. Merrill et al., Wyoming Gap Analysis Project. 33. Ibid.

Bright Green Hues Are Rare

225

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 226

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 227

11. Biological (Cryptobiotic) Soil Crusts of the Red Desert Jack States

o the casual observer a soil crust appears to be a curled-up sheet or mound of variously textured dried mud. But should these structures fail to disintegrate with the addition of water, one will quickly realize that they are much more than dirt. Biological crusts often appear as smooth, fragile, waferthin aggregations of surface soil particles, or they are variously usually dark colored, roughened mounds that can be up to fifteen centimeters tall. Sometimes referred to as cryptobiotic crusts (“crypto“ means “hidden”), they are communities of microscopic organisms — predominantly cyanobacteria (blue-green algae), green algae, filamentous fungi, lichens, and mosses — found within water-stable aggregations of silt and clay particles held together by the bodies and adhesive properties of the colonizing organisms. Biological soil crusts (bsc) are present at the surface of all but sandy desert soils. Apart from human-caused disturbances, the rate at which crusts develop is strongly influenced by climatic factors, especially drought, and by soil conditions. As measured by observations of recovery following disturbance, the time required for soil crusts to reach maturity ranges widely from two years in cool deserts to over three hundred years in extreme deserts.

T

Ecological Benefits Perhaps the greatest asset of bsc to the fragile habitats in which they occur is their reduction of the erosive effects of wind and water. They also make nutrients, especially nitrogen, available to plants. Consequent enhancement of seed germination and plant growth indirectly contributes to the long-term stability of arid and semiarid landscapes. bsc are now regarded as important components of healthy arid and semiarid deserts, where there is diverse plant cover and little or no evidence of active erosion. Red Desert Biological Soil Crusts A recent survey showed bsc to be widely but sparsely distributed in the Red Desert.1 Well-developed, mostly continuous crust cover was documented only in areas free from local grazing. One can most readily view them in the cushion plant communities on rock outcrops and escarpments that experience extreme weather conditions. Cold winters, low precipitation, strong winds, and the absence of grass and shrub competition promote both cushion plants and bsc formation. Well-developed bsc can also been seen

227

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 228

inside properly maintained Bureau of Land Management exclosures, where the effects of livestock trampling and grazing have been eliminated. Because the exclosures allow for impacts of wildlife (antelope, rabbits, ground-dwelling rodents), it is clearly evident that crust development in the past has not been severely affected under natural environmental conditions. Ancient, mature crusts were discovered as discontinuous, fragmented patches harbored beneath sagebrush canopies, a distribution indicative of a long history of disturbance.

preservation, additional and repeated soil disturbance may preclude the possibility of restoring the crust community to a predisturbance condition.

Note 1. Jack S. States, Commentary on Biological (Microphytic) Soil Crusts in the Rawlins Resource Management Area (Project Report, Biodiversity Conservation Alliance, Laramie, Wyo., 2004).

Impacts and Cumulative Effects (Trends) Crust communities are exceptionally vulnerable to surface disturbance and have a slow recovery time. Usually fifty to two hundred years or more are needed to restore the functionality of their original condition. The physical pounding by seismic equipment, euphemistically dubbed “thumper trucks,” has disturbed thousands of acres of the Red Desert in the search for coal-bed methane deposits. In an alarming report, the extractive minerals industry estimated that by 2015 there will be a massive grid of gas wells taking up 55,000 acres for well pads, and thousands of miles of roads and pipeline rights-of-way. These installations have the potential to destroy bsc throughout the Red Desert ecosystem, thereby exposing millions of tons of topsoil to the erosive forces of wind and water. Although the Bureau of Land Management is mandated to protect and manage threatened and sensitive plant species, the ecological importance and potential destruction of “cryptos” have been either inadvertently neglected or conveniently ignored in land use plans and environmental analyses. As one range manager commented, “If you can’t see ’em, they ain’t there.” Far from their “esoteric” and “intangible” labels, crust microorganisms are a vibrant and dynamic part of natural cool-desert ecosystems. In the absence of human-related impacts they have thrived for thousands of years, providing ecological stability and balance in extreme environments. In the absence of conservation and 228

Jack States

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 229

human history

Proulx_05_pp.171-230:Layout 1

8/4/08

6:37 PM

Page 230

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 231

12. Early People of the Red Desert Dudley Gardner

Paleo Period The Paleoindian period marks the oldest occupation of southwest Wyoming. It is thought that humans first lived here sometime before 12,000 bp during the transitional period that marked the end of the last ice age.1 Archaeological evidence suggests that they specialized in hunting megafauna, primarily large bison. Recent excavations conducted at the Pine Springs site southwest of Green River suggest that although camel, horse, and bison bones are present, no evidence of human activity is associated with these remains.2 Environmental changes led to the loss of forage for the large mammals, and as grazing areas diminished, many of the large fauna either disappeared or became smaller in size. The human hunters of the area had to adjust to the loss of mammoth and large bison, which disappeared as savanna-like conditions gave way to a drier environment and low grasses and sage came to dominate the region. During much of the region’s prehistoric past, hunters killed large game with spears tipped with beautifully crafted projectile points. One of the distinguishing characteristics of the Paleoindian period is the quality of the stone projectile points. Fluted Clovis and Folsom points reflect some of the most advanced lithic manufac-

turing techniques in the world. The Lancelot Eden point with parallel flaking and a diamond-shaped cross section reflect the skill of individuals whose intimate knowledge of stone shows in the tools. The subsistence and settlement patterns during the Paleo period are not clearly defined in southwest Wyoming. Some researchers believe that Paleoindian groups practiced a foraging strategy that targeted large mammals, with little investment in specific food storage places.3 Others hold that big game was not as important as once thought and that animals like jackrabbits were hunted extensively. These researchers think that Paleoindians collected grasses and stored foods in anticipation of winter. Evidence of small-game procurement is seen at several sites. At the Blue Point site and at 48sw8842 in the Washakie basin, at around 9500 to 9300 bp, a variety of small and medium-sized mammal remains were found, suggesting that at least some of the Paleoindian groups relied on a variety of mammals for food. Johnson and Pastor suggest that, by 9500 bp at Blue Point, inhabitants of the interior basin were using many resources, and the investigators have moved back the date of the Paleo/Archaic transition. With the exception of the Finley site (ca. 8800 –9000 bp) near present Farson, the sites excavated reflect this Paleo environmental and cultural strategy.4

231

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 232

Archaic Period Around 8500 bp there was a dramatic warming and drying climate change that affected the way these people lived. Phases of the Archaic period span the years from about 8500 to 2000 bp. In general, mammals became smaller, a change apparently related to a loss of grass and browse. The warmer temperatures reflected global changes and a reduction in precipitation as the world’s deserts expanded. In southwestern Wyoming it meant heavier dependence on smaller game and plant resources. The Archaic period has been subdivided into broad categories called the Early and Late Archaic.5 The Early period spans 8500 to 4300 bp. Archaeologist Michael Metcalf divides the Early period into the Great Divide Phase (8500 – 6500 bp) and the Opal Phase (6500 – 4300 bp). The Great Divide Phase begins with the onset of the warming and drying period. The archaeological record indicates that residents of southwest Wyoming began to live in pit houses. Relatively shallow, these houses had floors excavated in sand or loose soil, a wooden superstructure, and hide covers. The leather sides of the structure may have had soil thrown up against the bottom edges to keep out wind and moisture.6 The exact date when these pit houses first appeared is not known. The discovery of the structures changed the archaeologists’ view of the Archaic residents. No longer were the inhabitants seen as just “nomads.” They seemed people with sophisticated strategies for acquiring and storing food. During the Opal Phase, seed-grinding evidence shows up increasingly. More seeds are found in cultural deposits. There are also more bison bones. Pit houses become more prevalent and are basin-shaped, with charcoal-stained fill. They range from two to five meters in diameter and average about thirty centimeters in depth.7 Most of the pit houses contain internal features, including hearths and storage pits.8 There is a great deal of variability in the size and nature of the pit house. This suggests two things: the length of time the house was occupied, and the size of the group 232

living in the structure. The size differences may also reflect when they were occupied. Smaller, more open structures may have been used in the summer, and larger, more deeply excavated closed structures in the winter.9 Although not many pit houses have been excavated dating to the Great Divide Phase, pit houses characterize the Opal Phase.10 The house pits so far excavated in southwestern Wyoming cluster at around 6000 to 4500 bp and 1800 to 1200 bp. These date clusters suggest that, in the Early Archaic and Late Prehistoric periods, pit house structures were likely a successful adaptation to the environment. The Late Archaic seems connected with a different strategy.11 A decrease in radiocarbon dates between 4600 and 4300 bp shows up in the transition from Early to Late Archaic.12 While the decrease in radiocarbon dates alone may not be significant enough to prove the transition, the change in the environment and faunal assemblages does reflect increased moisture in the region. The Late Archaic spans the period from 4300 to between circa 2000 and 1800 bp. The era is marked by what archaeologist Steve Creasman calls a decrease in dependence on plant resources and an increase in large animal use.13 Although researchers found a larger number of bison bones in the faunal assemblage during this period, antelope, jackrabbit, and other small game remained important food sources. Seed processing, common in the Late Prehistoric period, is evident, but that evidence does not appear consistently.14 The Late Archaic is split into two phases. The period from 4300 to 2800 bp has been classified as the Pine Springs Phase. From 2800 bp to about 2000 bp is called the Deadman Wash Phase. Possibly during the Late Archaic there was an increasing reliance on larger mammals for food. There was still a high dependence on small animals for protein, but the archaeological record suggests a small change in the type of animals hunted.15 The evidence at present is inconclusive as to whether or not there was a transition toward hunting larger mammals. As more archaeological excavations are conducted, a clearer picture of the Late Archaic should emerge.

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 233

Late Prehistoric The Late Prehistoric period begins about 2000 to 1800 bp. The period is divided into three phases and reflects a time of rapid change and probably increased population in the region. Metcalf divides the early portion of this period into the Uinta Phase (2000 – 650 bp) and the Firehole Phase (650 –250 bp).16 Within the Late Prehistoric period stands a distinctive culture—the Fremont. The culture’s name comes from a river in central Utah that was in turn named after the nineteenth-century explorer John C. Fremont. The Fremont Phase spans the time from 2000 to 500 bp.17 The Fremont people grew corn and lived on the southern fringes of the Wyoming Basin. The Uinta, Firehole, and Fremont phases illustrate the complexity and diversity of cultures that evolved in the Wyoming Basin. The Protohistoric Phase is the most recent, beginning with the arrival of Europeans about 250 bp and ending with the removal of the Utes and Shoshonis to reservations in the late nineteenth century. The Late Prehistoric period is complex. To simplify, the Wyoming Basin can be divided roughly along the present Colorado-Wyoming border. But this is not a neat division. North of the border, the Uinta Phase marks the arrival of pottery to the region. Three types of pottery have been identified. Creasman proposed “two pottery traditions consisting of three styles.” The earliest tradition (ad 650 –1300) is a Fremont-like style called Black Buttes Gray. The next type postdates ad 1300 and is made up of Boar’s Tusk and Skull Point Gray wares. These two types have been loosely associated with the movement of Shoshonean and/or Crow groups into the area. Added to these three pottery types are isolated ancient Pueblo peoples’ potsherds discovered in southwestern Wyoming.18 There is debate as to when the bow and arrow arrived in southwest Wyoming. Archaeologist George Frison postulates that it arrived during the Late Prehistoric period on the northwest Plains. This period, Frison contends, “is recognized in the archaeological record by a change in projectile point types and sizes that most

investigators think is the result of the introduction of the bow and arrow.”19 The bow and arrow did aid solitary hunters. The archer had better accuracy than a spear thrower when shooting up and down hills. Manifestations of the Fremont culture can be seen in the Wyoming Basin in the form of rock images, ceramics, and maize. Fremont rock images show distinctive trapezoidal human forms, snakes, and mountain sheep. The Fremont people also made a grayware pottery tempered with sorted sand or limestone. Corncobs and even storage granaries have been located along the fringes of the southern Wyoming Basin. The Fremont stored corn in Brown’s Hole. Two cobs in the park returned dates of 870 ± 60 bp and 700 ± 30 bp.20 A variety of granaries were constructed in the area. Some were made from stone, and others from wood. Stone granaries vary in construction. Some were beehive shaped, others were built up against a cliff face using the cliff as the back wall, and still others were built up between fallen rocks. In all cases the stone granaries were protected from rain and snow by a rock overhang. Mud or adobe granaries have also been found. These had wooden frames and were covered with clay. In these granaries the wood was lashed together with string made from moistened bark or root, and the whole structure then covered with mud. The Fremont people often lashed wood beams together with bark or root twine. Evidence for lashing has been discovered in several places. At one location the granary frame was made of wood pieces lashed together and covered with a mud seal; corn was stored inside the granary.21 In almost all cases the granary had an entry that was covered with a flat rock or grinding stone. During the Uinta Phase there was a reliance on seeds, but antelope, bison, and deer are also present in the archaeological record. There seems to have been a more intense system of pronghorn and bison procurement during this phase.22 Outside groups seem to have also entered the region at the time, including the Fremont people, who had a significant presence in the area. At the Wardell site, Plains groups may have been present. George Fri-

Early People of the Red Desert

233

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 234

son recovered pottery and Avonlea projectile points at Wardell similar to those found on the northwestern plains produced by migrating Athabaskan groups.23 During the Late Prehistoric period, bison kill sites are most numerous. The Bessie Bottom, Woodruff, and the Barnes site, along with the Wardell site, all show large bison kills.24 Not only were the large bison kill sites and antelope kill sites in the Prehistoric period, but the Eden-Farson site, a Prehistoric/Protohistoric site dating to AD 1720 ± 100, contained evidence of 212 pronghorn, and Firehole Basin No. 11, dating to 625 ± 50 bp, contained 37 pronghorn.25 Communal antelope hunting seems likely at both of these sites.26 Along the Green River and its tributaries in Brown’s Hole and in present-day Dinosaur National Monument we find the arrival of maize agriculture. Fewer granaries with corncobs have been found in Brown’s Hole that date to the Firehole Phase (650 –250 bp), suggesting there may have been a decline in the region’s population.27 Too little is known about this period to determine whether there was an actual population decline or simply a falloff in the number of sites of the excavated period. By the end of the Late Prehistoric, Shoshonean groups, including the Utes, Bannocks, and Shoshonis, were well established in the Wyoming Basin. They were the people living in the region when the Europeans arrived. Protohistoric Period and the Arrival of Europeans Around 1700, Shoshonean groups that probably included the Yamparika Comanches dominated the Wyoming Basin. Their cousins, the Bannocks, hunted the western extremes of the region while the Utes lived in the southern Wyoming Basin—in a possible state of uneasy truces that led them to either get along with or fight the Comanches. The Utes seemed to have more tranquil relations with the Shoshonis. According to Domínguez and Escalante, the Utes claimed that the Yamparika Comanches lived north of the Uinta Mountains. The Utes, Sioux, Crows, and Black234

feet raided the region to the south for horses, but right up to their removal, in the Protohistoric and Historic periods, it appears that Shoshonean (or Numic) speakers controlled the Wyoming Basin. Traditionally, the Utes and the Shoshonis were divided into bands. The White River and Uintah Utes lived along the southern fringes of the Wyoming Basin. The Uintahs would cross the Uinta Mountains to hunt buffalo; the White River Utes would hunt deer in Middle Park and North Park, and buffalo in the vast expanses between the Yampa and Sweetwater rivers, either when they had the permission of the Shoshonis or when they felt strong enough to venture north on their own. Domínguez and Escalante, who traveled just south of the Uinta Mountains in 1776, noted that the Utes had a dreaded fear or respect of the Yamparika Comanches, who lived north and east of where they traveled.28 The Yamparika Comanches were directly related to the Shoshonis. The link between the two groups is most clearly seen in their linguistic ties.29 Finding food consumed so much of the Shoshonis’ time that many of their stories and tales described their lives and world in terms of hunting and gathering.30 They hunted a variety of animals and gathered seeds and plants in an area that to some seems arid and barren. They were so efficient at this that when Europeans arrived, the Shoshonis had surplus foods, furs, and hides to trade. The Shoshonis became wealthy traders with influence far beyond the boundaries of their nation. The Shoshonis in the area were actually the Eastern Shoshonis.31 Why they were called the Shoshoni is not known. As Virginia Cole Trenholm and Maurine Carley point out, “their Arapaho name, E-wu-ha-wu-si (People-Who-Use-Grass-or-Barkfor-Their-Lodges), probably comes nearer their tribal sign—a serpentine gesture — in reality the in-and-out motion used in weaving their shelters.”32 The term “Eastern Shoshoni” reflects where this group lived more than it reflects tribal affiliation. As anthropologist Ake Hultkrantz pointed out, the Eastern Shoshonis consisted primarily of the Buffalo Hunters, or Kucundika (“of the prairies”); the Sheep Eaters, or Tukudika (“in the mountains”); and the Dove Eaters.33 And while there are currently numerous

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 235

Shoshoni tribes in the Great Basin, the Eastern Shoshonis are commonly known as “the Shoshonis.” It was the Dove Eaters whose homes centered around Fort Bridger. Once they acquired the horse, Shoshoni culture changed, and they acquired the horse because of the arrival of the Spaniards far to the south.

Notes 1. “bp” stands for “before present.” 2. Robert L. Kelly, Early Holocene Prehistory in the Intermontane West: Reinvestigation of the Pine Springs Site, Southwestern Wyoming, NSF 9973272, Final Report (Laramie: University of Wyoming, Department of Anthropology, 2001), 54. 3. David E. Johnson and Jana Pastor, eds., The Blue Point Site: Paleoindian/Archaic Transition in Southwestern Wyoming (Rock Springs, Wyo.: Western Archaeological Services, 2003), 18. For the cultural chronology of northeastern Utah that borders the southwestern section of the Wyoming Basin, see Clay Johnson and Byron Loosle, Prehistoric Uinta Mountain Occupations, Ashley National Forest Heritage Report 2-02 (Vernal, Utah, 2002). The authors divide the area’s prehistory into the Paleoindian period (ca. 13,500 – 8500 bp), the Archaic era (8500 –2000 bp), the Late Prehistoric period (2000 –175 bp), and the Historic era, which begins with Domínguez and Escalante’s arrival in the Uinta basin in 1776 (pp. 13–16). They place the Fremont occupation of the area in the Late Prehistoric period and date this phase to the period spanning 2000 to 650 bp. 4. Kelly Pool provides an excellent discussion of rabbit consumption in the Wyoming Basin from about 9360 to 1730 bp. See Kelly Pool, “Eating Roots and Rabbits for 7600 Years in Southwest Wyoming: Results of Data Recovery at 48SW8842” (paper presented at the Colorado Council of Professional Archaeologists 2001, La Junta, Colorado, March 2001; manuscript on file with Metcalf Archaeological Consultants, Eagle, Colo.), 1– 7. Johnson and Pastor, Blue Point Site; Howard Haspel and George C. Frison, “The Finley Site Bison Bone,” in George C. Frison and Larry Todd, eds., The Horner Site (Orlando: Academic Press, 1987), 475– 491. 5. Michael D. Metcalf, “Contributions to the Prehistoric Chronology of the Wyoming Basin,” in Perspectives on Archaeological Resources Management in the Great Plains, ed. A. J. Osborn and R. C. Hassler (Omaha, Neb.: I&O Publishing, 1987), 233–261. 6. For the nature of pit houses in southwest Wyoming, see Joni Stainbrook, “Open Trench Inspection and Testing Results for the Mountain Gas Resources, Lizard Head 3-28 Pipeline, Sublette County, Wyoming” (Rock Springs, Wyo.: Western Archaeological Services, 1999); Lynn L. Harrell and Scott T. McKern, “Maxon Ranch: Archaic and Late Prehistoric Habitation in Southwest Wyoming” (Rock Springs: Archaeological Services of Western Wyoming College, 1986); Lynn L. Harrell, Ted Hoe-

Early People of the Red Desert

235

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 236

fer III, and Scott T. McKern, “Archaic Pit Houses in the Wyoming Basin,” in Changing Perspectives of the Archaic on the Northwestern Plains and Rocky Mountains, ed. Mary L. Larson and Julie E. Francis (Vermillion: University of South Dakota Press, 1997), 335–367; and David E. Johnson, “Archaeological Investigations at 48SU261 Sublette County, Wyoming” (Rock Springs, Wyo.: Western Archaeological Services, 2005), 60. Note: Archaeological Services of Western Wyoming College is now Western Archaeological Services, located at Western Wyoming Community College in Rock Springs. 7. Ted Hoefer, “Archaeological Investigations of the Bald Knob Site” (Rock Springs: Archaeological Services of Western Wyoming College, 1988). 8. Johnson and Pastor, Blue Point Site, 19. 9. Ibid., 19 –20. For a discussion of pit house size variability and function, see P. A. Gilman, “Architecture as Artifact: Pit Structures and Pueblos in the American Southwest,” American Anthropologist 52 (1987): 538 –564. 10. Johnson and Pastor, Blue Point Site, 19. For further discussions on pit houses, or basin houses, see William Lane Shields, “Basin Houses in Colorado and Wyoming: Delineation of a Culture Area and Parsing Hunter-Gatherer Modeling” (master’s thesis, University of Colorado, 1998); Cheryl Harrison and Steven D. Creasman, “Archaeological Investigations along the Eden Valley Landfill Access Road, Sweetwater County, Wyoming” (Rock Springs: Archaeological Services of Western Wyoming College, 1989); and Matthew D. Kautzman, “Archaeological Data Recovery at the Moxa Housepit Site (48LN616) along the Williams Field Services Mapco North Lateral Loop Pipeline Lincoln County, Wyoming” (Rock Springs, Wyo.: Western Archaeological Services, 2005). 11. Johnson and Pastor, Blue Point Site, 19 –20. 12. Ibid., 19. The Late Archaic period is divided into the Pine Spring Phase (4300 –2800 bp) and the Deadman Wash Phase (2800 –1800 bp); Metcalf, “Contributions to the Prehistoric Chronology,” 245. 13. Steve Creasman, “The Altithermal: Paleoenvironmental Reconstructions and Subsistence Change in Southwestern Wyoming,” in Perspectives on Archaeological Resource Management, ed. Osborn and Hassler, 283–297. 14. Johnson and Pastor, Blue Point Site, 20. For a discussion of lack of antelope procurement at Trappers Point during this period, see Mark E. Miller, Paul H. Sanders, and Julie E. Francis, eds., The Trappers’ Point Site (48SU1006): Early Archaic Adaptations in the Upper Green River Basin Wyoming (Laramie: Office of the Wyoming State Archaeologist, 1999), 1:21. 236

15. Johnson, “Archaeological Investigations at 48SU261,” 87–124. 16. Metcalf, “Contributions to the Prehistoric Chronology.” 17. Cf. Johnson and Loosle, Prehistoric Uinta Mountain Occupations, 14–16; and Alan P. Reed and Michael D. Metcalf, Colorado Prehistory: A Context for the Northern Colorado River Basin (Denver: Colorado Council of Professional Archaeologists, 1999), 6. 18. Johnson and Pastor, Blue Point Site, 20. 19. George C. Frison, Prehistoric Hunters of the High Plains, 2nd ed. (New York: Academic Press, 1991), 111. 20. The 700 ± 30 bp date is the corrected ams (accelerator mass spectrometry) date (SR-5994); A. Dudley Gardner, Heidi Guy Hays, Todd Kohler, Krista Snyder, Barbara Clarke, and Kevin W. Thompson, “Buster Flats Class III Cultural Resource Inventory Moffat County, Colorado” (Craig, Colo.: Bureau of Land Management Little Snake Resource Area, 2002), 138. The date of 870 ± 60 bp is from Beta 196824. 21. A. Dudley Gardner, Gabriele Elliott, and Melissa Pola, “Granaries in the Douglas Creek Drainage Basin” (paper presented at the Annual Colorado Council of Professional Archaeologists Meeting, Grand Junction, Colorado, March 2005; manuscript on file Western Wyoming Community College, Rock Springs); A. Dudley Gardner and Barbara Clarke, “The Fremont and Plant Resources along the Colorado-Wyoming Border,” Wyoming Archaeologist 46, no. 1 (Spring 2002): 5–13. 22. See Kevin Thompson and Jana Pastor, People of the Sage: 10,000 Years of Occupation in Southwestern Wyoming (Rock Springs: Archaeological Services of Western Wyoming College, 1995); and Craig Smith, “Seeds, Weeds, and Prehistoric Hunters and Gatherers: The Plant Macrofossil Evidence from Southwestern Wyoming,” Plains Anthropologist 33 (1989): 141–158. 23. Johnson and Pastor, Blue Point Site, 21. George Frison, “The Wardell Buffalo Trap (48SU301): Communal Procurement in the Upper Green River Basin, Wyoming,” Anthropological Papers of the Museum of Anthropology, University of Michigan 48 (Ann Arbor, 1973). 24. Frison, “Wardell Buffalo Trap”; Scott McKern, “Archaeological Investigations at the Bessie Bottom Site (48UT1186): Late Prehistoric Animal Procurement in the Upper Bear River Drainage” (Rock Springs: Archaeological Services of Western Wyoming College, 1988); Scott McKern, “The Barnes Site (48LN350): A Late Prehistoric Bison Jump in the Western Green River Basin” (Rock Springs: Archaeological Services of Western Wyoming College, 1995); W. F. Shields, “The Woodruff Bison Kill,” University of Utah Anthropological Papers 99, Misc. Paper 21 (Salt Lake City, 1978). 25. Miller et al., Trappers’ Point, 296 –297. The dates for Firehole Basin

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 237

No. 11 are 525 ± 135 bp. See Patrick Lubinski, “Pronghorn Intensification in the Wyoming Basin: A Study of Mortality Patterns and Prehistoric Hunting Strategies” (Ph.D. diss., Department of Anthropology, University of Wisconsin, Madison). 26. Alan D. Reed, ed., Kern River 2003 Expansion Project Wyoming: Cultural Resources Mitigation Report (Montrose, Colo.: Alpine Archaeological Consultants, 2005), 14. 27. Johnson and Pastor, Blue Point Site, 21. 28. Fray Angelico Chavez, trans., The Domínguez-Escalante Journal: Their Expedition through Colorado, Utah, Arizona, and New Mexico in 1776 (Salt Lake City: University of Utah Press, 1995). The possibility of the presence of Comanches in the region is noted in Johnson and Loosle, Prehistoric Uinta Mountain Occupations, 16. 29. Ernest Wallace and E. Adamson Hoebel, The Comanches: Lords of the South Plains (Norman: University of Oklahoma Press, 1952). 30. Anne M. Smith, Shoshoni Tales (Salt Lake City: University of Utah Press, 1993), 107. 31. There are three classical ethnographers of the Shoshonis: Robert H. Lowie, who published his findings in 1909 and 1924; Julian H. Steward, whose works were released in 1937 and 1938; and D. B. Shimkin, whose work came out in 1947. Lowie’s work consists of “The Northern Shoshoni,” Anthropological Paper of the American Museum of Natural History 2 (1909), 165–306; and “Notes on Shoshonean Ethnography,” Anthropological Papers of the American Museum of Natural History 20 (1924), 185–315. Julian H. Steward published extensively on the Great Basin tribal groups. His classic works are “Linguistic Distributions and Political Groups of the Great Basin Shoshoneans,” American Anthropologist n.s. (1937): 625– 634; “Basin-Plateau Aboriginal Social Political Groups: Analysis of Data,” Bureau of American Ethnology Bulletin 120 (1937): 230 –258; and “The Foundation of Basin-Plateau Shoshonean Society,” in Language and Cultures of Western North America: Essays in Honor of Sven S. Liljebad, ed. Earl H. Swanson Jr. (Pocatello: Idaho State University Press, 1970), 113–151. All of the above have been compiled by David Hurst Thomas’s in A Great Basin Shoshonean Source Book (New York: Garland, 1986). Left out of Thomas’ Source Book is D. B. Shimkin’s Wind River Shoshoni Ethnogeography, Anthropology Records 5 (Berkeley: University of California, 1947), 4. Since Shimkin’s whole focus is the Wind River Shoshonis, his work is essential to understanding the ethnohistory of this group of Native Americans. Shimkin and Lowie provide the best ethnographic data about the Eastern Shoshonis. Their work is closest in time to when the Shoshonis were placed on the reservation and when they were no longer allowed to follow traditional procurement

and trading cycles off the reservation. For a more recent work on Great Basin ethnography, see Warren L. D’Azevedo, ed., Handbook of North American Indians: Great Basin, vol. 2 (Washington, D.C.: Smithsonian Institution, 1986), in which Shimkin discusses the Eastern Shoshonis (308 –335). 32. Virginia Cole Trenholm and Maurine Carley, The Shoshoni: Sentinels of the Rockies (Norman: University of Oklahoma Press, 1976), 3– 4. 33. David Agee Horr, ed. American Indian Ethnohistory: California and Basin-Plateau Indians (New York: Garland, 1974), 11; Annual Report to the Commissioner of Indian Affairs, 1849 –1850 (Washington, D.C.: Gideon and Co., 1850), 1002 –1003. Mid-nineteenth-century tribal boundaries and a basic description of the Shoshonis are provided in a letter from John Wilson to the commissioner of Indian Affairs. Writing from Fort Bridger on August 22, 1849, Wilson provided the following report: “Their claim on boundary is to the east from Red Buttes on the North Fork of the Platte, to its head in the south, across the mountains, over to the Yan pa pu [Yampa River], till it enters Green or Colorado river; and then across to the back bone, or ridge of mountains called the Bear River Mountains, running nearly due west towards the Salt Lake, so as to take in most of the Salt Lake, and thence on to the Sinks of Mary’s or Humboldt’s river; thence north to the fisheries on the Snake river, in Oregon [present Idaho]; and thence south [their northern boundary] to the Red Buttes, including the source of Green river—a territory probably three hundred miles square, most of which has too high an elevation ever to be useful for cultivation of any sort. In most of these mountains and valleys it freezes every night in the year, and is in summer quite warm at noon and to half-past three o’clock p.m. Nothing whatever will grow of grain or vegetables; but the most [abundant plants] are nutritious grasses grown with the greatest luxuriance and the valleys are the richest of meadows.” John Wilson divided the Shoshonis into two groups. These divisions are relatively simplistic and show that he did not have a clear understanding of the Shoshonis. Nonetheless his description is somewhat enlightening regarding the conditions of the Shoshonis in 1849: “Among the Shoshoni there are only two bands, properly speaking. The principal or better portion are called Shoshoni, or Snakes, who are rich enough to own horses; the others, the Sho-sho-coes, or Walkers are those who cannot or do not own horses.” He adds: “Both bands number probably over one thousand lodges of four persons each. Of the relative portion of each band no definite account can be given; for, so soon as a Sho-sho-nie becomes too poor to, or does not, own a horse he is at once called a Sho-shocoe; but as soon as a Sho-sho-coe can or does own a horse, he is again a riding Indian, and therefore a Sho-sho-nie.”

Early People of the Red Desert

237

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 238

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 239

13. The Shoshonis and Westward-Bound Emigrants Dudley Gardner

gainst a backdrop of well-developed Native American trading patterns, westward-bound American travelers began to interact with the Shoshonis in the early 1840s. The Shoshonis, already skilled at trading with other tribes and fur trappers, expected to profit from and retain their homeland through the increasing emigrant trade. The years from 1840 to 1860 would determine whether they would be successful. In 1841 John Bidwell, John Bartleson, and a company of pioneers set out for the west coast. Leaving Westport, Missouri, on May 18, they knew their destination lay to the west but did not have a clear idea of the route. On the way they joined up with the first westering party of Catholic missionaries, which included Father DeSmet. Although mountain man Thomas Fitzpatrick served as their guide as far as Idaho, if it had not been for friendly Indians helping these two diverse groups, they might not have survived. In his reprint of the Bidwell diaries, Doyce B. Nunis calls this party “the first planned overland emigration west to California.”1 Although others had made the overland journey to the Pacific slope via South Pass before 1841, their destination had been Oregon. Further south, beginning in 1829, California-bound trading caravans had followed the Santa Fe Trail and the Old Spanish Trail or other variants to reach the west coast.

A

In his work California, John W. Caughey sees the BidwellBartleson group as “the entering wedge for the new type of migration to California.” The emigrants were bent on living permanently in what was to become the Golden State.2 However, they had to trace a vague, untraveled route from the Bear River to California. Earlier, Jedediah S. Smith had reached California in 1826 and 1827, and Joseph R. Walker had made the journey in 1833, but their reports of the way west were sketchy. There were no detailed maps or guides to tell the traveler where water could be found.3 Setting out into the unknown, they were adventurers with a dream, but no precise road map.4 Reverend Joseph Williams, who was with the BidwellBartleson party when it crossed the Green River in 1841, noted that the emigrants actively sought trade with the Shoshonis in southwestern Wyoming. On July 11, not far from Little Sandy Creek, west of South Pass, Williams wrote: “Today we lay by for the arrival of the Snake Indians to come and trade for articles, and a man was sent to tell them to come.”5 Thirteen days later on the Bear River, the party once again “rested and waited for the Snake Indians to come and trade with us.”6 Trade was essential since their animals had worn down and food supplies had been depleted. James John, who also accompanied this 1841 emigrant train, stated that on the Green River the travelers traded “with the 239

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 240

Indians and Trappers for horses and buffalo robes.” He also noted that the trappers and Shoshonis camped with them for several days as they traveled west.7 In The Plains Across, John Unruh wrote that Bidwell and Bartleson “employed and even kidnapped Indians to travel with them as guides, pointing out the trail and identifying good grass and water locations.” Hugh Skinner, traveling along the Hastings Cutoff in 1850, acknowledged if it hadn’t been for Shoshoni Indians directing them to water, they would never have found it. As Skinner also observed, “Many overlanders willingly entrusted their stock, wagons, belongings and even families to Indian swimmers and boatmen at dangerous river crossings all along the trail.” J. M. Shively explicitly stated in his 1846 guidebook that “you must hire an Indian pilot [when] you [are] at the crossings of the Snake River, it being dangerous if not perfectly understood.”8 The Shoshonis assisted Captain John Charles Fremont in his highly publicized 1843 exploration of southwestern Wyoming and the surrounding area. When Fremont traveled along the Bear River valley, he sought out a Shoshoni village for the purpose of trading for horses. He wrote, “We purchased eight horses, for which we gave in exchange blankets, red and blue cloth, beads, knives, and tobacco, and the usual other articles of Indian traffic.” He did not reveal how many knives or blankets he gave in the exchange, but judging from the experiences of others, the cost was relatively high. Fremont also obtained from the Shoshonis “a considerable quantity of berries of different kinds, among which service berries were the most abundant; and several kinds of roots and seeds, which we could eat with pleasure, as any kind of vegetable food was gratifying to us.”9 Fremont’s remarks provide excellent insight into the problems westward-bound travelers faced. Although meat abounded and the emigrants often carried flour, they soon ran out of foods containing vitamins A and C. Scurvy became a dreaded possibility. Passing through areas that contained nutrition-rich plants and roots, emigrants and travelers from the east did not recognize the available food sources, nor did they know how to process them. In other areas some emi240

grants picked berries they found along the trail, leaving few for those that followed. For essential vegetables and fruits they often depended on trading with local Indians. Edwin Bryant, in 1846, went west on the back of a mule. The former editor of the Louisville Courier, he provided a revealing description of his encounter with Native Americans in present southwestern Wyoming. On July 18 he recorded: “Several Indians visited our camp in parties of three or four at a time.” Showing he did not quite comprehend their actions, he stated, “An old man and two boys sat down near the door of our tent, this morning, and there remained without speaking, but watchful of every movement for three or four hours. When dinner was over, we gave them some bread and meat, and they departed without uttering a word.” The Indians were not begging but simply following cultural norms of polite visitation, and since apparently they could not speak English, and Bryant could not communicate to them in their native tongue, silence seemed appropriate. Bryant appreciated the setting he was in, as he wrote, “Circles of white-tented wagons may now be seen in every direction, and the smoke from the camp-fire is curling upwards, morning, noon, and evening.” A large “number of oxen and horses are scattered over the entire valley grazing upon green grass. Parties of Indians, hunters, and emigrants are galloping to and fro, and the scene is one of almost holiday liveliness.” He added, “It is difficult to realize that we are in a wilderness, a thousand miles from civilization. I noticed the lupin, and a bright scarlet flower, in bloom.”10 By the time emigrants began to travel across southwest Wyoming, native horse herds provided fresh animals to wagon trains. James John in July 1841 traded with “Snake Indians and trappers for packhorses” along the Green River.11 John Bidwell noted on July 25, 1841, that “a good gun is worth as much as a horse.”12 The Shoshonis, as well as former trappers (or “hunters,” as they were called in the emigrant journals), profited handsomely from trading with emigrants. At times Fort Bridger was essentially run by Native Americans. One traveler noted that the

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 241

Shoshonis greeted the emigrants and were prepared to trade goods. On July 9, 1846, William E. Taylor wrote that “16 miles Braught us to [Fort] Bridger Shoshoni in abundane.”13 The next day the wagon train “Lay By, Indians visited us in great numbers.”14 Joel Palmer provided yet another insight into Fort Bridger’s operation when he noted the post was “generally abandoned . . . during the winter months.”15 Since emigrant traffic virtually ceased from September to May due to the harsh winters, it made little sense to live in the high mountain valley all winter. Following tribal traditions that had allowed the Shoshonis to thrive in this high-elevation steppe, the trappers, traders, and their wives also migrated to sheltered sites better suited for enduring the longest season in the Rocky Mountains. While crossing the Green River in 1849, James A. Pritchard, bound from Kentucky to California, noted an encampment of Shoshonis along the river. He wrote in his diary that the “Indians . . . had a great number of fine horses. And for which they asked a big price.” He identified them as “Shoshoni or Snake Indians.” “They are,” according to Pritchard, “decidedly the best looking and most intelligent Indians that I ever saw. They possess an affability and suavity of manners not common to the Red Men of the Forest. Their Women are handsome delicate and genteel looking.”16 That same year, another traveler commented on the number of horses the Shoshonis owned. Lewis Shutterly, who was traveling west, wrote that he “crossed Thompson’s fork of green river along this is a beautiful valley and there is an Indian camp of about 200 wigwams and 1500 to 2000 Indians of the snake tribe they appear a harmless people and own many fine horses in which they take much pride they being their sole property they are remarkable good riders.”17 At Fort Bridger on July 25, 1849, James Wilkins noted that “there are here 20 or 30 families of mountaineers principally Canadian French married to Indian women, and living in tents of skins.” A keen observer, he added that “considerable white frost was on the ground this morn. . . . Altho’ there is plenty of grass and fine water, a beautiful looking

trout stream close by they say they cannot raise any vegetables on account of the coldness of the nights.”18 Initially unable to grow vegetables, the residents were dependent on Native Americans for these food sources. The newcomers’ use of wild plants, likely harvested by Native Americans, is borne out by the presence of Indian ricegrass found in one of the storage areas within Bridger’s trading post.19 Writing of his trip west in 1850, Byron McKinstry, an emigrant bound for California, stated that, at the crossing of Ham’s Fork, “Indians are plenty they are a better looking race of Indians than I have ever seen before. They are all of the Snake tribe.” Noting that the Indians had been greatly enriched by their trade with white travelers, he described them in this manner: “They are whiter, better formed, better dressed, more intelligent and [own] more property than their brothers the Pawnees, Omahaws, Otoes, and etc. near the frontiers.” Mounted on splendid horses, the men and women “gracefully [rode] about as if they were Lords and Ladies with nothing else to do,” McKinstry continued. “Many are dressed in the European stile, probably procured from the Emigrants or picked up.” Claiming that the Shoshonis were excellent traders, he explained “they are willing to trade guns, clothes, and etc., but their horses they could not be induced to part with.”20 Atop the Bear River Divide in western Wyoming in 1851, P. V. Crawford wrote, “Here the scenery is grand. The Green River Valley in the east and the Bear River Mountains on the west.” Headed for the Willamette Valley in Oregon, this emigrant from Indiana noted: “Good grass and pure springs all the way. We met a lot of Indians today. They had been out on a hunting expedition, had plenty of game and were in good plight and good humor. Here everything is most lovely.”21 The terms used in Crawford’s diary show that he and his companions did not fear the Shoshonis. In 1851 Susan Amelia Cranston noted that along the Green River “Indians and white men” had horses and oxen for sale. She added, “Turned our stock loose and did not guard them.”22 Other women commented that “Indians and white men” bartered and traded freely. Mary Louisa Black, who took the Overland Trail to Fort

The Shoshonis and Westward-Bound Emigrants

241

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 242

Bridger on her way west to Oregon, recorded in her 1865 diary, “Aug. 2 Last night ice froze 1/8 of an inch in thickness. The Snake Indians came in to barter fish and antelope hides for bread, coffee.”23 The Shoshonis consistently provided travelers with items that aided them in their trip west.

242

Notes 1. Doyce B. Nunis, ed., The Bidwell-Bartleson Party 1841 California Emigrant Adventure: The Documents and Memoirs of the Overland Pioneers (Santa Cruz, Calif.: Tanager Press, 1991), 1. 2. In ibid., 8. John W. Caughey, ed., California (Englewood Cliffs, N.J., 1959), 213. 3. Nunis, Bidwell-Bartleson Party, 8. The Bidwell diary points out some of these concerns after the travelers were well into present Nevada. A very few maps did exist for the west when Bidwell and Bartleson traveled to California. Some had been made by Spaniards; others, by Jedediah Smith. These maps, however, were not easily consulted, and they showed general features and not necessarily where springs and good feed for horses could be found. See Dale L. Morgan, Carl I. Wheat, and Jedediah Smith, Jedediah Smith and His Maps of the American West (San Francisco: California Historical Society, 1954). 4. Nunis, Bidwell-Bartleson Party, 42, 43, 47, 48, 49, and 50. On August 26, 1841, Bidwell writes: “Our course intersected an Indian Trail, which we followed directly north towards the mountains, knowing in these dry counties the Indian Trails always led to the nearest water.” On August 28, 1841, he adds: “A Shoshoni Indian came to our camp [not too far from the Great Salt Lake]; from him we learned that there were more Indians not far off who had horses. Several men and myself went in search of them. Having gone about 5 miles, up hills and down hills covered with thick groves of cedar (red), we unexpectedly came to an Indian, who was in the act of taking care of some meat-venison—which he had just killed; about half of which we readily purchased for 12 cartridges of powder and ball. With him as pilot we went in pursuit of other Indians; he led us far up in the mountains by a difficult path, where we found two or three families, hid as it were from all the world, by the roughness of nature. . . . Not succeeding finding horses, we returned to the camp.” Bidwell’s entry of September 3, 1841, states, “Four or 5 Indians came to camp—bought three horses off them,” and the next day’s entry notes, “Bought a few service berries off the Indians.” By October, Bidwell and “Captain B.” (Bartleson) were worried that they had failed to find the allimportant Saint Mary’s River, which was supposed to take them to the Sierra Madres of present California. Now deep into Nevada, they were among the Western Shoshonis and had mixed experiences with the Native Americans. On October 6 they hired an Indian “to pilot us on.” On October 7, Bidwell writes: “The Indian pilot remained with us—the river spread into a high, wide swamp, covered with high cane grass—Indians

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 243

are numerous. Encamped by the swamp about dark, having come about 25 miles —water bad — no fuel, excepting weeds and dry cane grass which the Indians had cut in large heaps to procure sugar from the honey dew with which it was covered.” Bidwell was served well by this pilot. Bartleson had briefly parted company with Bidwell and his guide and nearly starved. Bidwell fared somewhat better, until the pilot left on October 16. Other Indians had assisted Bartleson, who, Bidwell notes, came into camp on October 16, “rather in a hungry condition, and had been traveling several miles without provision, excepting a few nuts which they had purchased from the Indians and which they had eaten on a very small allowance.” Having lost their guide, they were now desperate. On October 23, Bidwell states that, “having no meat than would last us 3 days, it was necessary to use all possible exertions to kill game, which was exceedingly scarce.” On the same day, an Indian boy ventured into camp and sold Bidwell acorns. Despairing and desperate, the BidwellBartleson party began to lose horses to the Indians, who “eat them.” By late October they entered into a new land, and the Native Americans were not as helpful—and were even hostile—to their party. They reached California in early November, but without Indian help to guide them across the desert, the question of whether they would have made it at all bears asking. 5. Ibid., 245. 6. Ibid., 246. 7. Ibid., 169. 8. John D. Unruh, The Plains Across: The Overland Emigrants and the Trans-Mississippi West, 1840 – 60 (Urbana: University of Illinois Press, 1979), 157. 9. Donald Jackson and Mary Lee Spence, eds., The Expeditions of John Charles Frémont, vol. 1, Travel from 1838 to 1844 (Urbana: University of Illinois Press, 1970), 475. Fremont makes several interesting points, including the name of the Green River: “This is the emigrant road to Oregon which bears much to the southward to avoid the mountains about the western head of Green River—the Rio Verde of the Spaniards.” He also notes that the Indian name for the Green is Seeds-Ke-Dee-Agie (466). 10. Edwin Bryant, What I Saw in California; Being the Journal of a Tour by the Emigrant Route and South Pass of the Rocky Mountains, across the Continent of North America, the Great Basin, and through California, in the Years 1846, 1847 (reprint, Minneapolis: Ross and Hainer, 1967), 144–145. 11. Nunis, Bidwell Bartleson Party. 12. Ibid., 37. Bidwell notes that “a cap lock [gun] is preferred, caps worth $1 per box.”

13. Within quotations, the original punctuation, spelling, and capitalization are used, such as this misspelling of “abundance.” 14. Dale Morgan, Overland in 1846: Diaries and Letters of the CaliforniaOregon Trail (Georgetown, Calif.: Talisman Press, 1963), 127. William E. Taylor was bound for California via the Central Overland Route. 15. Ibid., 35. 16. Ibid., 97. John Unruh, in The Plains Across, 164, writes: “One wagon outfit, preparing to pack in to California from the Sweetwater River, swapped a wagon, its contents, and four yoke of cattle to Snake Indians for five ponies.” 17. Lewis Shutterly, The Diary of Lewis Shutterly (Saratoga, Wyo.: Saratoga Historical and Cultural Association, 1981), 26. Normally, the number used to estimate population by lodges is four per lodge. Since 4 ∑ 200 = 800, Shutterly’s estimate might be too high. Even at a higher rate of six Shoshonis per lodge, Shutterly’s count may still be too high. 18. John Francis McDermott, ed., An Artist on the Overland Trail— 1849 Diary and Sketching [of ] James F. Wilkins (San Mateo, Calif.: Huntington Library, 1968), 57. 19. Linda Scott Cummings, Paleo Research Laboratories, Golden, Colo., pers. comm., August, 1992. A. Dudley Gardner and David E. Johnson, “Results of the 1992 Archaeological Excavations at Fort Bridger” (Rock Springs: Western Archaeological Services, Western Wyoming Community College, in preparation). 20. Bruce L. McKinstry, The California Gold Rush Diary of Bryon N. McKinstry 1850–1852 (Glendale, Calif.: Arthur H. Clark, 1975), 193. 21. P. V. Crawford, “Journal of a Trip across the Plains, 1851,” Quarterly of the Oregon Historical Society 25 (June 1924): 147. “P. V. Crawford” is apparently Philemon Vawter Crawford, who settled in the Willamette Valley in 1851. 22. Kenneth L. Holmes, ed., Covered Wagon Women: Diaries and Letters from the Western Trails, 1840–1890, vol. 3 (Glendale, Calif.: Arthur H. Clark, 1984), 114. 23. Bert Webber, ed., The Oregon and Overland Trail Diary of Mary Louisa Black in 1865 (Medford, Oreg.: Pacific Northwest Books, 1989), 45.

The Shoshonis and Westward-Bound Emigrants

243

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 244

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 245

14. An Anthropological Impression of Rock Art in the Greater Red Desert Russel L. Tanner

he Rocky Mountain west and nearby portions of the Great Plains and Great Basin hold abundant rock art features that represent a diversity of ethnic influences, functions, and ages. Rock art, including petroglyphs incised or ground into rock surfaces and pictographs painted onto rock, are among Wyoming’s most important historical and archaeological resources. For many years most archaeologists tended to gloss over the significance of rock art sites because there were no precise ways to date them that would facilitate comparison with other artifacts and features. In the past two to three decades it has become clear that rock art can be dated by a number of means, although some remain controversial. Perhaps the most effective dating method is still the old standby of relative comparison of artifact types and styles. Comparisons of five hundred years of the historical record in the form of diaries and journals of early American travelers and European explorers with hide paintings, ledger art, and other renditions allow contextual analysis of the cultures that likely created the etchings on stone. If the relatively new chemical dating techniques, of which there are several, prove reliable, then the framework for analysis may be broadened, perhaps covering nearly the entire range of human presence in the New World. Rock art is now recognized

T

worldwide as one of the most information-laden archaeological features from which to make anthropological inference about past human behavior. During the last twenty years there has been important research involving assumptions and approaches in understanding rock art, especially in the work of Jim Keyser and a number of others (Keyser and Klassen 2001; Keyser and Poetschat 2005), including several local Wyoming investigators (Tanner 1990, 1991, 2002; Tanner and Vlcek 1995). Such researchers have been evaluating the historic and cultural significance of ceremonial and biographical rock art—most created in the past five hundred years. This approach, while not entirely new, has recently been applied with significant success in the study of ethnohistory of indigenous peoples the world over. The second approach involves much older rock art sites, especially in Europe and Asia but also in Africa, North America, and Australia, demonstrating the development of what may be a near-universal Paleolithic shamanistic ceremonial realm (Clottes and Lewis-Williams 1996). David Whitley, who has worked extensively in California and also studied rock art in Africa, believes that this practice developed into the vision quest and other traditions among many Native North American cultures (Whitley 2000). This is especially intriguing in light of the very early date of 11,500 years before present (bp), obtained from the Black Rock 245

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 246

petroglyph panel in the Red Desert (Tanner et al. 1995; Whitley et al. 1996). While the biographical and ceremonial styles are by far predominant in the greater Red Desert area, there are two sites with very significant shamanistic aspects. Although both sites probably also have biographical information, the strong spiritual presence and meaning of the most prominent images seem undeniable. Black Rock may be one of the earliest rock art sites on the continent; and the other site—Tolar—is probably evidence of quite recent cultural interaction among resident Red Desert cultures and trading/raiding interlopers from the south and perhaps the southeast. This essay is a collection of impressions of several clusters of rock art sites within the greater Red Desert area. The following discussion consists of thoughts about the ages and origins of several sites, a general description of some of the main images, and exploratory ideas about their meanings, contexts, and circumstances. The Site Clusters Rock art sites in the greater Red Desert region are clustered in three places—Boar’s Tusk, Seedskadee, and East Flaming Gorge. One seemingly unassociated site (Black Rock) is worthy of special mention. In the Boar’s Tusk region north of Rock Springs there are a number of biographical petroglyph sites from the historic period, or perhaps a little earlier. All of the sites in the region are incised, although there are a few pecked-in images, particularly at Cedar Canyon. White Mountain has an abundance of bison and elk images as well as horse-mounted warriors and one place perhaps representing a birthing chair, or midwife’s station. On a patinated surface in one part of the site are totem figures, including bear paws and turtles, while on the opposite side of a narrow alcove there is a soft sandstone rock, with deeply embedded handholds. About a 246

hundred feet to the east along the rock escarpment is an area where baby-footprint images are ground into the rock patina. Recently White Mountain has suffered increased vandalism, including severe graffiti in the area of the “birthing chair.” One panel actually collapsed in 1994. The Pine Canyon site is on state land and depicts mountain sheep and early historic-era human figures armed with flintlocks, seemingly battling each other. No horses are depicted at the Pine Canyon site, which seems curious. This might indicate that White Mountain is the latest of the big sites in the Boar’s Tusk region. Unfortunately, Pine Canyon suffered several forms of vandalism in the past, including the actual cutting away of some panels and blue paint applied on others. The third well-known site in this area, Cedar Canyon, has incised geometric designs and mountain sheep but no bison or horses. There are also many enigmatic images that seem to represent some type of plant branches, perhaps tobacco, although the renderings are not like other representations presumed to be Indian tobacco. Local archaeologists refer to these forms as “spinies,” because of their stickery appearance. Cedar Canyon is the only site where these images are represented. Geometric designs and mountain sheep are the most common features at Cedar Canyon. Two sites not in the Boar’s Tusk area have similar-aged biographical and perhaps ceremonial images. Tolar is a somewhat isolated site several miles southeast of the Boar’s Tusk cluster but is probably of a similar age. This remarkable ceremonial and biographical site is postulated to have been a Comanche horse-trader location, based on a detailed image of a mounted warrior with a unique headdress. Powder Mountain contains mostly charcoal pictographs, relatively recent, with many horse images. A few petroglyphs, perhaps older than the pictographs, are also known in the Powder Mountain area. These obviously historic images may have been drawn by reservation-era Ute or Shoshone Indians. A number of other archaeological features in this area include a large corral or extensive fence made of felled, interwoven juniper

Russel L. Tanner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 247

branches, as well as several circular stone alignments that may have been guard posts. The second cluster, Seedskadee, is a separate, seemingly discrete group of rock art sites in the LaBarge–Big Piney area, about a hundred miles north of the town of Green River. The sites seem to be clustered within easy reach of the Green River itself, and some are right along the river’s edge. These sites are technically not within the Red Desert, though culturally they are undoubtedly related in many ways to sites within the desert proper, and any consideration of rock art in the region should include them. The impressive LaBarge Bluffs site images indicate a story line that may be based on actual events, including battle scenes and the capture of a female by horse-mounted raiders, followed by a scene of two dance lines, which might represent an initiation ceremony. Two railroad locomotives, and a male profile drawing with the notation “Dan 1877,” all seem to date much of this site to the post–transcontinental railroad era. Elsewhere at the site an exquisite elk with heart line and ribs may date somewhat earlier, as may several other images that do not seem associated with the story line sequence. There are a number of apparent Native American petroglyphs at the Names Hill site. The Native images include horses and tepees. Names Hill has long been known for its hundreds of American pioneer inscriptions, but the Indian petroglyphs have not been previously documented. However, pioneer J. Goldsborough Bruff (1873) drew some petroglyphs he observed in the 1850s that may or may not be those known at the site today. Bracketed ages documented by Jim Keyser and George Poetschat indicate that the site was occupied by Indians during the same years when the numerous inscriptions were made by American sojourners at this major crossing of the Green River on the Sublette Cutoff of the California-Oregon-Mormon pioneer trails. A rock art site called Gateway is located on the west side of the Green River near the confluence with Fontenelle Creek. The images at Gateway seem to be somewhat older than those at LaBarge Bluffs. The Gateway images include a battle scene in

which two groups use spear throwers to launch atlatl darts at each other. The panel also has some ceremonial aspects such as a possible pipe, as well as other human figures. The third cluster of rock art sites is in the Flaming Gorge area, to the east of the man-made lake. Undoubtedly a number of sites are under the reservoir and were not recorded or studied before the Flaming Gorge Dam was constructed in the early 1960s. These sites are situated along major tributaries of the Green River, including Salt Wells Creek, something of a northern boundary for the type of rock art styles represented in this cluster. The Sugarloaf site is a single, though very complex panel of images that include perhaps Fremont-style animals—mountain sheep, elk, coyote, and a possible cat. Richards Gap, along Red Creek, has a black painted image that is too faint to be discernible, as well as a beautiful red painted handprint. The two painted symbols are located on the back wall of a small alcove that features a protruding slab of rock containing nine basin metates. The Richards Gap images, like those at all of the sites in this cluster, have a Fremont feel to them. Along Salt Wells Creek is a group of sites that are typically reminiscent of Fremont-style human and animal figures, although they generally are not as large nor as refined as the rock art figures in the core Fremont culture area around Vernal, Utah. Similar, though somewhat more classically Fremont panels are found in Irish Canyon in Colorado just south of the Wyoming border, and also at Minnies Gap on the Wyoming-Utah state line east of Flaming Gorge Reservoir. Finally, there is a unique site, Black Rock, a single pecked art panel situated north of Point of Rocks. This site seems to be the oldest rock art site in the region. The pecked images of mountain sheep, and perhaps other animals, together with human figures and possibly even some representations of nets or trapping devices for game procurement, have in many places been repatinated after having first been etched into the original patina of the stone surface. This repatination process would seem to have required thousands of years to have occurred as a natural phenom-

An Anthropological Impression of Rock Art

247

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 248

enon. Black Rock, more than any other site in the Red Desert, seems to have shamanistic representations perhaps along the lines of those described by Whitley in the Great Basin (Tanner et al. 1995; Whitley et al. 1996).

dle Park region of the Colorado Rockies. In this way the Red Desert is connected to the agricultural cultures located on the lower Platte and Missouri rivers. Trade

Geography The setting in which culture groups operate is an important consideration in discussion of the anthropological issues of Native cultures in any region. It is apparent that rock art styles and images under discussion here are related to the geography of the Red Desert. It seems likely that the unique rock spire known as Boar’s Tusk, together with the flattop mesas (Pilot Butte and North and South Table mountains) and the mobile landscape of the Killpecker Sand Dunes in the central Red Desert, attracted attention from aboriginal people living in the area. This unique landscape could hardly have gone unnoticed by the inhabitants. Another critical issue of geography involves the major watercourses of the region—in this case the Green River. This region, as characterized by the Palisades near the city of Green River and the Firehole region are actually the beginning of the famed red rock landscape that continues hundreds of miles into the Four Corners area of the southwest. Prehistoric people used waterways as guides, and river valleys provided relatively easy corridors for the movement of people from one region into another. The Green River thus became the gateway to the intensively occupied southwest. Indeed, the Rocky Mountains of Wyoming are the “top of the west” in terms of being the source of all the major river drainages in the western United States—the Yellowstone to the Missouri, the Platte to the Missouri, and the Green River to the Colorado. This places the Red Desert at a critical juncture, where the Rockies could be crossed relatively easily. From South Pass the Sweetwater River provides a route to the North Platte of the Missouri, while the North Platte proper provides something of an eastern boundary of the Red Desert, coming out of the North Park–Mid248

Trade and the associated exchange of ideas have long been known to have been critical in the development of cultural complexity of the upper Green River basin (Ewers 1954). It seems reasonable to expect that trade networks are reflected in the rock art of this region. Obviously, the trade item exemplifying the historic contact period in this region is the horse, which is everywhere in the desert rock art. From what are apparently the earliest stick-figure horses at White Mountain to the equestrian warriors at LaBarge Bluffs and the rider with the ornate headdress at Tolar, horses are probably the most typical item in rock art of the greater Red Desert. With the actual trade items also came ideas and cultural traditions from surrounding regions. Anthropologist John Ewers (1954) postulated the presence of a trade center of considerable time depth in the upper Green River basin. This ancient trade network likely foreshadowed and became the foundation for the Green River fur trade rendezvous of the early to mid-nineteenth century. Most of the rock art can, to a greater or lesser extent, be associated with this trade network. Warrior Societies: Horses and Guns in Rock Art As trade in the area expanded, so did the ethnic diversity of cultural traits (Secoy 1992) we now see reflected in the region’s rock art. Indeed, the Red Desert became a crossroads and marketplace of the Central Rocky Mountains. The resident people commonly called the Snakes were indigenous Shoshonean cultures who eventually diversified into the Wind River Shoshone, Fort Hall Bannock, and Fort Duchesne Ute peoples. They served as trading intermediaries with northern plains tribes such as the Crows and

Russel L. Tanner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 249

the Blackfeet, Columbia Plateau groups such as the Salishes and Nez Percé and Kootenais, as well as the Comanches, who were Shoshonean speakers in the southwest. Undoubtedly, over time there was also cultural contact among many other groups, including the Athapaskan speakers of the southwest (Apaches and Navajos), and Algonquin speakers from the Great Plains, the Arapahos and Cheyennes in particular. From these vast reaches a diffusion of cultural traits occurred in the greater Red Desert and especially along the Green River. Keyser and Poetschat (2005) have demonstrated what had been clear to many observers of Red Desert rock art for a long time: most of the rock art is heavily laden with imagery of the Plains Indian warrior tradition. Rock art images in the region then could reasonably be expected to be a blend of many of these diverse cultures. Thus the prominence of the bison, tepee, and horse images at many sites reflects many different groups. Depictions of firearms, noted especially at Pine Canyon and the LaBarge Bluffs, indicate that the earliest group to possess such weapons were the Blackfeet, but the images more likely point toward the somewhat later possession of guns by resident Shoshonean groups. The interesting presence of what local archaeologists call spinies might represent some kind of plant products traded into the area, even perhaps tobacco traded by Crow people. While objects like trade beads are not easily recognized in the rock art of the area, such items are very heavily represented in a number of protohistoric archaeological sites in the region. By placing rock art imagery into a chronological series, Keyser and Poetschat (2005), and to a lesser degree Tanner and Vlcek (1995), have demonstrated how rock art functioned as a pictorial representation of coup counting, horse raiding, and other issues prevalent in Plains warrior societies. Since the Green River basin—and indeed all of the Central Rocky Mountains—was the place where much of this historical action took place, rock art offers another avenue for investigation, together with journal accounts, ledger art, government documents, and other sources for study of this aspect of western American history. Like ledger art,

pictographs and petroglyphs offer an opportunity to see these issues from the Native peoples’ perspective. The people who inhabited the greater Red Desert over a millennia or more should be expected to have been a mixture of cultures, undoubtedly with a heavily Shoshonean foundation. However, many different cultures may have inhabited an area over time. Unfortunately, with the slight exception of rock art, most archaeological evidence reveals few linguistic associations. How a given ethnic group has manifested itself differently on the landscape and in the archaeological record at different times is not easily discerned. No doubt, somewhat distant Shoshonean groups, including Ute and Comanche horse traders, had great influence in the area. The mere diffusion of horses into the region transformed the pedestrian Shoshonean residents into mounted hunters and trading people and vastly expanded their scope of influence. Eventually, horses enabled the Shoshones to become part of the plains bison economy and escape the limitations of the Great Basin subsistence strategy sometimes referred to as the “desert culture.” Connections with Plains tribes eventually allowed the people who became the Shoshone alliance under Chief Washakie to control a vastly larger area reaching far onto the plains and, through cultural affinity, well into the Columbia Plateau. But even before the Shoshones became part of the Plains bison hunting tradition, they were likely a distinctive culture that may be reflected in some of the earliest rock art images of the region, including especially the pecked images at many sites, as well as the total array at the Black Rock site. A number of archaeological questions are only recently beginning to develop. What is the relationship of Late Prehistoric hunting and plant-gathering cultures to later cultural development via interpretation of the rock art in the Flaming Gorge area? How is the group of cultural traits commonly referred to as “Fremont” to the southwest of the greater Red Desert region related to cultures in the Red Desert, and especially, to the apparent representation of “Fremont” things in rock art sites in the Flaming

An Anthropological Impression of Rock Art

249

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 250

Gorge area (Loosle and Wilson 1998)? Are these sites diffused from contemporaneous village cultures of the southwest, or might they be older cultures that developed into the Fremont groups? Likewise, could the Fremont-like traits seen in Red Desert rock art rather represent the breakup of the village cultures and a return to traditional hunting-and-gathering lifeways more like the lives of Archaic Shoshonean residents of the upper Green River basin? Physical traits evident in a number of skeletal remains of ancient peoples found in the Red Desert have led University of Wyoming anthropologist George Gill (1991, 431– 447) to refer to a “Red Desert people.” The evidence, partly falling within measurement parameters for Shoshonean people, indicates they were probably not much divergent from other groups, especially some Athapaskan examples. So it might be that the people of the Red Desert proper were an admixture of several physically identifiable culture groups. In this case one can only guess at the language and other cultural traits they would have exhibited. It seems likely that these people held many old Great Basin traditions, perhaps evident in the historically known Sheepeater traditions of the mountains to the north (Loendorf and Stone 2006). There could also be reflections of Plains traditions, including the Sun Dance, as well as Fremont-derivative traditions in places. Regardless of how one might interpret rock art and other archaeological evidence from the greater Red Desert, it is readily apparent that cultural development, as reflected in the rock art, is complex and not easily pigeonholed into any single set of cultural associations. There is a vital need for land management agencies to pay much more attention to rock art resources in the Red Desert. Over the past twenty years, as vast portions the desert have become industrialized, the Bureau of Land Management and state landowners have paid far too little attention to the unique rock art sites, which are rich with cultural information. In some cases, like White Mountain and Tolar, agencies have documented the sites, but there is still virtually no protective management presence at the actual sites. This allows tremendous “use wear” by visitors 250

whose behavior is not controlled at the sites, resulting in actual vandalism of these wonderful and irreplaceable resources. Finally, one positive result of increased activity in the greater Red Desert has been the stronger involvement of Native Americans in management and interpretation of rock art. For this researcher the incredible value of a real human perspective not only on rock art but also on indigenous peoples’ lifeways has been enhanced by discussion and friendship over many years with numerous American Indians from many tribal affiliations and cultural positions within their tribes.

Russel L. Tanner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 251

References Bruff, J. Goldsborough. 1873. “Indian Engravings on the Face of the Rocks along Green River Valley in the Sierra Nevada.” Annual Report of the Smithsonian Board of Regents 107:409 – 412. Clottes, Jean, and David Lewis-Williams. 1996. The Shamans of Prehistory: Trance and Magic in the Painted Caves. Harry N. Abrams, New York. Ewers, John C. 1954. “The Indian Trade of the Upper Missouri before Lewis and Clark: An Interpretation.” Bulletin of the Missouri Historical Society 10:429 – 446. St. Louis. Francis, Julie E., and Lawrence L. Loendorf. 2002. Ancient Visions: Petroglyphs and Pictographs of the Wind River and Bighorn Country, Wyoming and Montana. University of Utah Press, Salt Lake City. Gill, George C. 1991. “Human Skeletal Remains on the Northwestern Plains.” In Prehistoric Hunters of the High Plains, by George C. Frison, 2nd. ed. Academic Press, San Diego. Keyser, James D. 2004. Art of the Warriors: Rock Art of the American Plains. University of Utah Press, Salt Lake City. Keyser, James D., and Michael A. Klassen. 2001. Plains Indian Rock Art. University of Washington Press, Seattle. Keyser, James D., and George Poetschat. 2005. Warrior Art of Wyoming’s Green River Basin: Biographic Petroglyphs along the Seedskadee. Oregon Archaeological Society, Publication 15. Portland. Loendorf, Lawrence L., and Linda Olson. 2003. “The Tolar Petroglyph Site.” American Indian Rock Art 29:1–10. Loendorf, Lawrence L., and Nancy M. Stone. 2006. Mountain Spirit: The Sheep Eater Indians of Yellowstone. University of Utah Press, Salt Lake City. Loosle, Byron, and Kelda Wilson. 1998. Vernal Area Rock Art. Published by Byron Loosle, Vernal, Utah. Secoy, Frank R. 1992. Changing Military Patterns of the Great Plains Indians. University of Nebraska Press, Lincoln. Tanner, Russel L. 1990. “Images from the Great Margin: Rock Art of the Green River Basin in Wyoming.” Paper presented at “Rock Art 90: A Symposium,” San Diego Museum of Man, San Diego. ———. 1991. “LaBarge Bluffs: An Important Protohistoric Rock Art Component along the Green River in Wyoming.” Paper presented at the 18th Annual American Rock Art Research Association Conference, Las Vegas, Nevada. ———. 2002. “Pictures by the Seedskadee: Cultural Implications of Two

Rock Art Sites along the Green River in Southwestern Wyoming.” Paper presented at the 67th Annual Meeting of the Society for American Archaeology, Denver. Tanner, Russel L., Joseph Bozovich, Julie E. Francis, and Ronald I. Dorn. 1995. “The Black Rock Petroglyph: A Possible Clovis-Age Rock Art Site.” Poster paper presented at the 53rd Annual Plains Anthropological Society Conference, Laramie. Tanner, Russel L., and David T. Vlcek. 1995. “Picturing the Past: Prehistoric and Historic Rock Art of the Green River Basin in Wyoming.” Paper presented at the Second Biennial Rocky Mountain Anthropological Conference, Steamboat Springs, Colorado. Whitley, David S. 2000. Art of the Shaman: Rock Art of California. University of Utah Press, Salt Lake City. Whitley, David S., Ronald I. Dorn, Julie Francis, Lawrence L. Loendorf, Thomas Holcomb, Russel Tanner, and Joseph Bozovich. 1996. “Recent Advances in Petroglyph Dating and Their Implications for the Pre-Clovis Occupation of North America.” Proceedings of the Society of California Archaeology 9:92 –103.

An Anthropological Impression of Rock Art

251

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 252

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 253

15. Traversing the Desert Annie Proulx

On the Trails Freighting, though less celebrated than covered wagon trains or the pony express, was an important part of the westward shift. All of the people scratching their way west needed supplies: bacon, nails, firearms, hardwood, wagon parts, forge iron, tools, flour, clothing, harness. Almost everything was made in the east and had to be hauled overland. Every spring freight wagon trains left the Missouri towns. A big freight wagon could carry five thousand to six thousand pounds of goods and used six yoke of oxen for power. The lead oxen were usually Texas longhorns, tough, well trained, and more intelligent than most cattle. A full train of twenty-five wagons needed three hundred oxen (with another thirty or so for spares), twenty-five bullwhackers singing their way along, as well as two wagon masters, a clerk, and a night herder.1 The teams moved slowly, about two miles an hour, but doggedly. In 1866 a twenty-three-year-old Civil War veteran and photographer-artist, William Henry Jackson, suffering simultaneously from a broken heart and itchy feet, hired on as a bullwhacker in a train bound for Montana. If you had no money, it was a way to get out west. Jackson wrote, “We bullwhackers were a curious, mixed crew. Mostly quite young, several of us ex-

soldiers, we had among us four or five middle-aged ‘professional’ men who dearly loved the hard life. . . . Several of the boys were of good address and obviously of some education; a few had been farmers in Pennsylvania, Ohio, Indiana and Illinois; four or five had been clerks in the cities . . . three of them were graduate roustabouts from the Missouri River school; one was a FrenchCanadian; and one . . . was a genuine jack-of-all-trades from Boston.”2 Jackson’s freighting trip took him over the California-OregonMormon Trail. The group had just passed Deer Creek telegraph station west of Fort Laramie when a mail carrier and several telegraph operators galloped past, yelling that Indians had attacked the station and set it on fire and that they were on the way to the next military post for help. By August the freighters reached the Sweetwater River, out of sugar, molasses all gone, the oxen sore-footed and limping. “Through all this country the going was very hard,” Jackson wrote. “Double teaming was the rule . . . and that cut our daily distances to a half or a third the usual mileage. But we knew that the Continental Divide was close at hand and whipped the panting bulls along. And on Monday, August 27, a great shout went up as the snow-capped Rockies suddenly loomed in the distance beyond a little ridge.”3 253

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 254

From the 1840s through the 1860s the emigrants, miners, government geographers, regular and volunteer soldiers, Civil War veterans, surveyors, telegraph workers and operators, traders, opportunists, scalawags, men of business, and adventure-seeking boys who crossed the country on the westering trails are often thought of today as restless Americans on the move. Yet many of them were foreign-born, perhaps to the extent that the westward expansion was as much a European as an American movement. Jackson, the flavor of the scene fresh in his mind, wrote: But that taken-for-granted opportunity, the wide continent quite reasonably bestowed on the citizens of the Republic, was something entirely new to the new people thronging in from Europe. It was sheer miracle. Free land? There had been no such thing in the old country for so many centuries that the very idea was fantastic—there was not even land that you could buy, except by accident or the very slow turnover of estates as family situations changed. And here in America was illimitable acres, and the kindly Homestead Act of 1862, which gave land to anyone who would take the trouble to farm it and live on it. The tide of European immigration gathered force and swept west with the restless Americans. You met every language and every racial type on the plains in the ’60’s and ’70’s.4 The original Howard Stansbury report of his 1849 –1850 expedition to the Great Salt Lake included a map drawn by Lieutenant J. W. Gunnison and Charles Preuss. Stansbury’s map shows his outbound route as the California-Oregon-Mormon Trail. But on the return he follows a more southerly Red Desert route along Muddy Creek and through Bridger Pass into the Sierra Madre/Medicine Bow ranges. The map is important, not only for its details of the Great Salt Lake area, but for reaffirming Bitter Creek and Sage Creek, and for adding Currant Creek, Quien Hornet Mountain, Bishop Mountain, and Bridger Pass to the topographical lore.5 254

The map also shows a section of the Cherokee Trail, “Evans route,” a little north of Stansbury’s route but paralleling it. And there is a faint dotted line across the Red Desert, for years the subject of much controversy. It is labeled “Jones Route 1850.” Who was Jones and what was he doing in this inhospitable country ominously labeled “artemisia barrens with some pasture on the water courses”? For decades no one knew. Carl Wheat, in his magisterial six-volume Mapping the Transmississippi West, wrote: “An ‘Evans Route,’ . . . had been opened by a party of 47 wagons from the Cherokee Nation in 1849. . . . A variant ‘Jones Route 1850’ is quite as interesting, little if anything being known about the party which traveled it.”6 Over the last two decades researchers have solved part of that mystery. The Cherokee Trail, roughly 1,300 miles long, became a major mid-nineteenth-century route for emigrants going to California. It was a tangle of links, shortcuts, and connecting routes to other trails. Today (2006) local stockmen, sheepherders, four-wheeldrive trippers, and energy company trucks continue to travel its stony backbone. In Wyoming one of the northern sections (which the Stansbury map shows as the “Evans route”) lay parallel to and a little south of the California-Oregon-Mormon Trail.7 In 1862 when Holladay’s Overland Stage line shifted south, it absorbed a northern section. The southern section, used by several 1849 and 1850 groups, was miles farther south. In the early 1980s the Wyoming State Historic Preservation Office began to systematically record trails, including the Cherokee.8 In 1994 an overview, largely based on material from the Cherokee Phoenix, documented the trail’s route through southwest Wyoming.9 Five years later another group of researchers produced an ambitious work describing the trail from Oklahoma to California.10 Since 2001, ongoing archeological surveys of trail remnants and fresh research continue to trace the confusing tangle of cutoffs and side routes.11 The Fletchers’ study, Cherokee Trail Diaries, concerns the mixed white and Cherokee companies of 1849 and 1850; they

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 255

used discovered travelers’ diaries to chart the routes. The 1849 group, the Evans/Cherokee party, was made up mostly of whites from around Fayetteville, Arkansas, as well as men of the Cherokee Nation. The whites apparently felt that having Cherokees in the party would protect them from hostile Plains Indians, but that did not mean they liked their traveling companions. When the Evans group finally separated at the “Independence Road” (California-Oregon-Mormon Trail) near Fort Bridger, one of the Pyeatts from Washington County, Arkansas (forced to relocate by the 1828 Indian boundary, and obviously holding a grudge), wrote: “[A] po[r]tion of those weakest retches was unwill[ing] to stay so the captain gave them leave to go on and we have not seen them since all and thankes be to God for thear departure for the bulk of them ar of the loast clas of beings that aint fit to liv or di. . . . they ar the cherikees and Fayettvill trash. . . . they come on to bridgers fort and thear many of them got drunk on whiskey at $1 apint and cut up tall shines.”12 The Evans company had planned to cover twenty miles a day. As much as possible they traveled on high ground, as from the ridges they could spy out landmarks, Indians, water, and grass. In many sections of the route they were trailblazers or, on known trails, the first travelers of the year. The company picked up the Santa Fe Trail in Kansas ahead of the hordes bound for the goldfields and remained fit. At Pueblo the company broke into three parts. The original plan had been to go west following the Canadian River across Oklahoma and the Texas panhandle, and on to Santa Fe. But reports of Indian agitation in that country persuaded them to change. They thought they could work their wagons up over the Rockies by tracing the headwaters of the Arkansas to its source, and then bear away northwest to Salt Lake City. Various Pueblo residents told them that it was impossible to do this with wagons. The only way through the high country was with a pack outfit. About thirty men of the party traded their wagons for mules. These packers hired a guide, the experienced mountain man Dick Owens. Evans and the main body decided to go north to St. Vrain’s

Fort, turn northwest to South Pass, and get on the CaliforniaOregon-Mormon Trail, then referred to as the Independence Road. They became the wagon party. And a dozen or so men turned around and headed back to Arkansas. The packers roughly paralleled today’s Route 287, past Steamboat Rock and Virginia Dale. Near today’s Tie Siding, Wyoming, they turned west. They crossed the Laramie River and, after twisting through the heavily forested Medicine Bow slopes above North Park, crossed the North Platte. Owens led them southwest to the Little Snake River, where, a generation later, settlers would create the hamlet of Savery. By way of Brown’s Hole and Fort Bridger, they arrived at Salt Lake City three weeks before Evans’s wagon party, having avoided most of the sagebrush ocean and the Red Desert badlands. The main Evans wagon party moved steadily northwest across the Laramie plains, around the south end of Elk Mountain, and through Pass Creek (so named by Stansbury the following year), the route of Ashley and Fremont before them. Although it was July, the nights were cold enough for morning frosts. They began to get sick with something many emigrants suffered—”mountain fever.”13 The Fletchers say that, apparently not finding enough grass and water, they did not go through Bridger Pass, as some researchers have believed. It was left to Stansbury, with Jim Bridger as guide, to be the first to record this pass in 1850.14 There was no road through Bridger Pass until 1858, when the army built one during the Mormon War. On July 18, near the present site of Rawlins, the wagon party turned sharply west on a route that would take them through the Red Desert basin.15 They had a hard time of it and left three oxen somewhere in the desert. Theirs was the first wagon train to cross the arid basin. Forty-nine wagons made a lasting impression on the fragile desert ground, and a year later Stansbury, traveling east from Salt Lake with Jim Bridger as guide, came again and again on Evans’s road. From the time the Evans party crossed the North Platte until

Traversing the Desert

255

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 256

they reached the main trail eighteen days later, they were stressed for water and grass. When they did find a little water, Pyeatt wrote: “We traveled on nex day the most of the men with out a drop of water till in the evning say one oclock all the men out huntin watter that had horses to ride. . . . those riding returning and find now [none] . . . when one of the men that was a foot found water. This water was from a heavy s[h]ower of rain that fell 2 days before and woeld not las more than 2 days longer had it not bin fer this water we must have suffered conssidrable for we did not find any more [water] till next night som 18 or 20 miles further.”16 Lieutenant Carrington, with Stansbury on the 1850 return trip, wrote that on September 21, as they left Miller Creek north of Bridger’s Gap and headed for the pass south of Medicine Butte (today’s Elk Mountain), they struck Evans’ road: “Bridger says it is very poorly grassed & watered—and will do only to pass very early in the season & even then it is a rugged & barren route.”17 Diarist Crawford, who rarely wrote more than a brief note on the date, mileage, and camp place, remarked with feeling: “From this Place Back 75 or 80 miles is a pore dreary, Rough and Hilly Country, destitute (almost) of grass, and no timber of any Kind except the Artemecia Shrub. . . . But 2 Springs have Been Seen in the last named distance; one Calebrate . . . the other Pure white Sulpher.”18 On August 6 the party reached “the Salt Lake Road East of Bridgger’s Fort on a Butifull Streem Called Blacks Fork of Green River.”19 From this point on, the journey was easier, as they were on a main trail. In 1850 alone, 65,000 people made the overland crossing to the goldfields of California.20 The Cherokee trail sections pioneered in 1849 attracted more Arkansas parties, among them the mysterious Jones, whose delicate trace curved across Stansbury’s map. The puzzle of who Jones might have been lies perhaps in the Cherokee Advocate. In the 1990s the Fletchers, reading through the paper, came across an article in the August 31, 1853, issue headlined “WILLIAM CLAUDE JONES’S ROUTE TO THE PACIFIC — 256

HIGHLY IMPORTANT STATEMENT.” These were the years when ideas for a transcontinental railroad sprouted like dandelions. The government throughout the 1850s actively encouraged searches for the best route. Explorers connected with the U.S. Topographical Corps looked for possible rail routes through the Rocky Mountains and the Sierras to California. Others traversing the country for their own ends took notice of easy grades and passes. Jones apparently was one of these railroad route scouts, and in the Cherokee Advocate article claimed he had found an easy route. He reported: Crossing the South fork [South Platte River] above St. Vrain’s Fort, the route will skirt the base of the mountains . . . and enter a pass where the South Fork of the Cache a la Poudre bursts from the Mountains. Then . . . through a prairie valley and over an elevated plain to the Medicine Bow mountain[s], near the source of Laramie river. This mountain is . . . covered with a dense forest of excellent timber for the construction of a road, some fifty miles southeast of the Medicine Bow Butte [Elk Mountain]. Emerging from this forest, it enters the North Park, passes the North Fork of the Nebraska or Platte, skirts the base of the main range [Sierra Madre], and crosses through a low prairie divide [Twin Groves] about two degrees south of South pass, on the head waters of the North Fork of the Yampa river [Little Snake River], and enters the desert valley of the Upper Colorado of the West [Green River]. . . . It will not be necessary to cut a single tunnel. Nature has graded the road. . . . The highest point of the pass is about 7,000 feet, and the ascent from the east is gradual.21 This was roughly the route that the packer contingent of Evans’s party had followed. Among the packers had been James Vann, who in 1853 returned to Arkansas and his position as editor of the Cherokee Advocate. He wrote an exposé of the article touting Jones’s “discovery”:

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 257

Cherokee Trail routes, 1849 –1850

Of a part of the route as described by Jones we had ocular demonstration. We crossed the South Platte at St. Vrain, entered the mountains where Cache la Poudre burst forth from its mountainous confines, bore a little north toward the Medicine Bow mountains, crossed the mountains into the Parks, passed over Laramie Plains to the headwaters of the Yampah river, down said river to the Vermilion mountains, across said mountains to Brown’s Hole, from thence to Bridgers: during our whole time in the mountains we did not see any of these ascents over which railroads can be built as if by magic. If this route is so far preferable to any other, it is strange that it is not the favorite route of Col. Benton, as it is one of Col. Fremont’s trails, as we were told so by our guide Owen, who was also Fremont’s guide along the same route. . . . we saw none of the fine timber as described . . . but we saw that which is not spoken of by Jones. We on the fifth of July saw and passed over . . . snow. If those mountains were passed . . . without being tunneled, many hundred miles would be added . . . by running around them.22

Jones might have been a member of one of the 1850 Cherokee/white parties that traveled overland from Arkansas to California. The route he described was similar to that taken by the 1850 Mitchell-Quesenbury group.23 There is, in fact, a reference to someone named Jones in Quesenbury’s diary entry for May 14, when travelers were still in Kansas.24 In April of 1850 the second wave of Arkansas white/Cherokee groups prepared to head out as four wagon groups and one packer group. The various parties decided it would be worthwhile to try new cutoffs from the Evans track. The main outfits were the Cane Hill California Emigrating Company, which had as members James Mitchell and William M. Quesenbury, both of whom kept diaries; Captain Edmonson’s ox train and wagon group; and Samuel Houston Mayes Cherokee group, with diarist John Lowery Brown. The most interesting person of any group was probably the twenty-seven-year-old Quesenbury, an educated man, a schoolteacher, an avid hunter and fisherman, something of an artist who made many sketches along the way, and a traveler who had survived journeys to Texas and Mexico. He apparently had a copy of

Traversing the Desert

257

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 258

Lewis Evans’s 1849 journal or a list of Evans’s mileages. Edmonson’s ox train, while still in Kansas, successfully hired guide Ben Simon for $61 and a mule to take them to Fort Bridger. Essentially the parties all followed the Evans route to Pueblo, where Quesenbury tried unsuccessfully to hire guide Charles Kinney.25 Here the packer party switched to mules, heaping their surplus possessions in a pile and setting fire to it to keep the overgreedy Pueblo residents from seizing their unwanted goods. When Edmonson’s ox train reached the Laramie plains, Ben Simon guided them west, away from Evans’s north trail. The parties behind followed Edmonson’s wheel ruts on what became the southern Cherokee Trail. They touched on the Little Snake River in the vicinity of present-day Baggs and continued west along the Cherokee rim through the southern Washakie basin. On June 28, at the confluence of Muddy Creek and Wild Cow Creek, Quesenbury and Mitchell struck out northwest, impatient at the slow rate of ox travel, hoping to discover a cutoff. Guide Ben Simon gave them as much information as he could. They set off up Muddy Creek, then bore left toward Mexican Flats and on to Barrel Springs Draw. They fought their way “through a sea of wild sage.” At night it was cold and they burned sage to keep warm. They skirted north of the striking geological formations called the Haystacks and avoided the badlands of Adobe Town and Skull and Prehistoric rims, though Quesenbury’s description seems to hint that they could see these landmarks. Their exact route is not known and, without discovery of additional letters and/or diaries, maddeningly unknowable. On July 1, Quesenbury wrote: “The country became more destitute today than yesterday. We rose very high ground after leaving the water. Pushed on, without any prospect of getting water, but in ardent hopes of it. The country was cut up with gullies; the soil loose, nothing but dusty sand and hard mud all cracked in millions of different ways. Clay bluffs sometimes heightened the miserable looking country. They were washed in various shapes. Sometimes a tower; sometimes pillars, steps, domes etc.” 258

Mitchell’s entry on the fourth day showed some asperity— no water, no grass, everyone hungry. They moved to “a Snowy mountain” a few miles distant where there was water, grass, and wood. On the fifth day, Quesenbury wrote: “We got up and saddled our animals as soon as it was light enough to do so. Struck out W, our old course. The country became no better as far as water was concerned. Passed some remarkable scenery of rocks but we were in no condition to enjoy such things. . . . Took off our packs. There is water below this place about a hundred yards, but Mr. Mitchell who tasted it thinks it a mixture of every thing vile,— alum, copperas, arsenic etc. He says it is too mean to drink, to taste or even to look at.” The next day, their sixth in the Red Desert, they reached Bitter Creek, which Mitchell described as “running west without grass or watter fit for use all crusted with Salt & Some poisonous matter.”26 But from a high hill, late in the evening, they could see the larger stream into which Bitter Creek emptied—the Green River, or, as Mitchell called it, “Mary river.” Their animals were weakened after the ordeal of dry travel, it was still a day’s hot, dry journey down to the Green, and then they had the problem of crossing it. But they were out of the Red Desert. General Rusling In 1866, anxious to cut the burdensome costs of maintaining scores of small posts and forts across the west, the army quartermaster general’s office sent a number of Civil War generals, now recast as inspectors general, to examine the posts from Fort Leavenworth to the Pacific. General James F. Rusling was one of these men, ordered to pay special attention to Salt Lake City and then to visit Washington, Oregon, Nevada, and Arizona.27 The general’s assessments would help determine which posts could be abandoned. In early October the general left Denver on the Overland Stage, headed for Salt Lake. He did not ride in one of the famous Concord coaches but traveled in a more lowly equipage, which he

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 259

described as “‘Red Rupert’ . . . a mountain mud-wagon, with a low canvas top, so as to be less liable to capsize in crossing the range.”28 The nights were cold, the mud wagon uncomfortable, the road a rocky mire, and the threat of Indians omnipresent. The coach made fairly good time —104 miles in twenty-four hours — the general dozing as best he could in the jolting wagon. At the North Platte they swapped Red Rupert for a light coach and continued on to Fort Halleck in Wyoming. The general was unaware they were crossing the continental divide until the driver mentioned it. The coach entered the Red Desert. Rusling wrote: Our first introduction to the Pacific slope was hardly an agreeable one. At our great elevation the night was bitterly cold, and we had shivered through long hours, in spite of our blankets and buffalo-robes. Routed out at 3 A.M., for breakfast, we straggled into the stage-station at Sulphur Springs, cold and cross, to find only dirty alkali water to wash in, and the roughest breakfast on the table we had seen yet, since leaving the States. Coffee plain, saleratusbiscuit hot, and salt pork fried—made up the charming variety, and we bolted it all, I fear, as surlily as bears. A confused recollection of cold, discomfort, and misery, is all that remains in my memory now of that wretched station at Sulphur Springs, and may I never see the like again!29 General Rusling was in no mood to appreciate the subtle beauty of the Red Desert, which he called “the Desert of the Mountains, the famous or infamous ‘Bitter Creek Country,’ accursed of all who cross the continent.”30 As they went on, one of the other passengers, apparently depressed by the landscape, “indulged too freely in Colorado whiskey” and went into delirium tremens. The general wrote: “He tried several times to jump out of the coach, and made the night hideous with his screams; but we succeeded finally in getting him down under one of the seats, and thus carried him safely along.”31

It was a bad night, for sometime later one of the coach’s thoroughbraces broke, and, as the general put it, “Then we had to go bumping along on the axle-tree for ten or twelve miles, until we reached the next station. This no doubt was a good antidote to John Barleycorn; but it scarcely improved our impressions of ‘Bitter Creek.’”32 On the Train The moment the transcontinental railroad was finished people rushed to take the trip. Most of these early tourists ardently desired to see wild Indians, a wish not often gratified. Young Robert Louis Stevenson (1850 –1894) traveled by train to California in 1879 against the advice of friends and family, for, even though he was pursuing the woman he loved, he was not well. Poor and ill, he rode in an emigrant car. He commented: “I saw no wild or independent Indian; indeed, I hear that such avoid the neighbourhood of the train; but now and again at way stations, a husband and wife and a few children, disgracefully dressed out with the sweepings of civilisation, came forth and stared upon the emigrants. The silent stoicism of their conduct, and the pathetic degradation of their appearance, would have touched any thinking creature, but my fellow-passengers danced and jested round them. . . . I was ashamed for the thing we call civilisation.”33 Stevenson longed to see the mountains, but when the train nosed into Wyoming’s Laramie Range, he was disappointed, and as they went on, depressed: All Sunday and Monday we traveled through these sad mountains, or over the main ridge of the Rockies, which is a fair match to them for misery of aspect. Hour after hour it was the same unhomely and unkindly world about our onward path; tumbled boulders, cliffs that drearily imitate the shape of monuments and fortifications—how drearily, how tamely, none can tell who has not seen them; not a tree, not a patch of sward, not one shapely or commanding mountain

Traversing the Desert

259

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 260

form; sage-brush, eternal sage-brush; over all the same weariful and gloomy colouring, grays warming into browns, grays darkening towards black; and for sole sign of life, here and there a few fleeing antelopes. . . . Except for the air, which was light and stimulating, there was not one good circumstance in that God-forsaken land.34 The evening the train left Laramie, Stevenson’s health cracked, and he spent the night walking up and down, longing for daylight. When it came, he saw they were still in the sagebrush: “That train was the one piece of life in all that deadly land; it was the one actor, the one spectacle fit to be observed in this paralysis of man and nature.”35 Roads across the Desert It is difficult for us to grasp the violent enthusiasm that accompanied the construction of the Lincoln Highway from Times Square to San Francisco. Americans celebrated with fireworks, fish fries, speeches, bonfires (which one supporter called “bond fires”), barbecues, and street dances. The oratory relentlessly drilled the word “progress” into people’s ears. Driving across the continent on the Lincoln Highway was a true adventure, “one of life’s big experiences,” as an Iowa newspaper editor said (an attitude that prevailed to the time of Jack Kerouac and even today). The Lincoln Highway was the country’s first coast-to-coast road, opening the American automotive frontier. For a while, it was the most famous road in the world. The Lincoln Highway was the brainchild of Carl G. Fisher, creator of the Indianapolis Motor Speedway, head of the PrestO-Lite Company. In 1912 he dreamed up the idea of a crosscontinent crushed-rock road of 3,389 miles, which he called the Coast-to-Coast Rock Highway. People believed it could be done. Had not the country built a railroad through the wilderness? Were they not building the Panama Canal after the French had given up? Was not everything 260

possible for Americans, the children of Progress? The media splash was immense. One of the biggest supporters of the idea was Henry B. Joy, the president of the Packard Motor Car Company. Joy put forward the name “Lincoln Highway” and worked desperately hard to see the idea come to life. By 1914 the Lincoln Highway Association had worked out a cross-country route of sorts and posted this loose connection of roads with blue-and-white signs declaring it the Lincoln Highway. It was, in fact, a barely traversable mess. Payson W. Spaulding of Evanston, the Wyoming attorney for the Union Pacific in addition to his regular legal work, accepted the appointment as chief consul of the Lincoln Highway for the State of Wyoming.36 It was up to the various states to choose the precise route through their lands.37 Immediately Spaulding found himself dealing with a hot controversy. In the miles between Laramie and Rawlins two towns, Medicine Bow and Elk Mountain, fought for the official route, each claiming the better road. Wyoming could not reach a decision on the rival routes. In December, A. R. Pardington, the association’s first secretary, wrote to Spaulding that the dispute “has got to a point where it is practically impossible. The actions of some of the people there are next door to criminal. . . . [N]owhere throughout the entire 3400 miles of length of the Lincoln Highway is there a condition anything like that between Laramie and Rawlins.”38 Medicine Bow had stolen a march on Elk Mountain by putting up unauthorized Lincoln Highway signs. The association eventually allowed both routes the coveted signs, leaving the choice up to the traveler. But in the summer of 1914, when the first tourists came driving through, the problem solved itself. Almost all the travelers chose the Medicine Bow route, as more than thirteen gates had to be opened and closed on the Elk Mountain road thanks to intransigent ranchers who balked at fencing off rangeland for the sake of a road.39 Medicine Bow went into the official guidebook as the Lincoln Highway route; Elk Mountain was mentioned as a side trip with many gates. The controversy was over.

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 261

For the Red Desert section west of Rawlins, John Hay of Rock Springs wrote Spaulding in November of 1914 that he would “arrange to have the distance between Green River and Wamsutter covered by Friday, November 27, providing we don’t have a storm so that it is impossible to make the trip.”40 That section of road needed more than signs. Generally, the roads remained frightful. In Wyoming, horseshoe nails punctured tire after tire. Another problem was bridges. In 1919 the association’s field secretary, Harry Ostermann, and a convoy of fifty-six military trucks and other heavy vehicles set out on a test of the rough roadway to see if it could take the weight. It could not. In sixty-five days the vehicles fell through more than a hundred bridges, almost half of them in Wyoming, where severe wind and aridity weakened bridge timbers. One of the soldiers on this trip was Lieutenant Colonel Dwight D. Eisenhower, later to make his presidential mark as a major road builder. Just west of Cheyenne they broke through their first Wyoming bridge. As they went on, they found stacks of heavy timber next to each bridge, courtesy of the Wyoming Highway Department. The morning after a big trout feast in Rawlins the convoy headed farther into the Red Desert. Colonel Charles McClure described this section of highway as “mountain desert of most desolate and monot[on]ous character” enlivened by “dry air wind and dust hardship continuous.”41 They covered fifty-eight miles in eleven discouraging hours, shoring up bridges all the way, finally camping at Tipton Station, empty of humans except for a few railroad workers; they endured a cold night when the temperature sank into the forties. The next day was awful, with high winds and severe dust storms that made it difficult to see the faint track: “The rutted wheel paths wound up and down through sand holes and washouts in wild passes and canyons. Some of the time, wrote McClure, they saw ‘magnificent palisades of stratified earth and rock rising to heights of seven hundred feet.’ Dust permitting, the view otherwise was of ‘utter desolation strewn with bones of animals/the intensely dry air absence of trees and green vegetation and parched appearances of landscape exert depressing in-

fluence on personnel.’ It was, he noted wearily, ‘altogether most tedious day of expedition.’”42 The Lincoln Highway literature contains scores of travelers’ stories of mishaps and trials through the Red Desert. Those who tried to shortcut across alkali flats sank to their running boards. In Walcott a family whose car broke down spent the night in the house of some residents who were not home. The road passed Como Lake, which Frank Leslie, riding the Union Pacific in his private car in 1877, described as a “little sheet of water . . . fed by warm springs . . . [with] wild ducks and the constant presence of an agreeable species of lizard, ranging from six to eighteen inches in length, which supplied the chief diet of the ducks.”43 In 1999 Gregory M. Franzwa described the same water as “the unappealing algae-covered Como Lake.”44 In 1921 the Townsend Bill became the Federal Highway Act. The famous highway itself vanished. The rigidly correct director of the new Bureau of Public Roads was Thomas Harris MacDonald, a man who lived by the rules. In 1925 the bureau did away with the Lincoln Highway, substituting a system of numbers.45 When Henry Joy died, his widow, Helen Joy, had a monument put up to honor his devotion to the Lincoln Highway. It was erected in the Red Desert southwest of Rawlins, “the most touchingly appropriate of places,” as a chronicler of the highway put it. “Here, for sixty-two years, a dark gray obelisk stood on the windswept earth. An inscription read, ‘That there should be a Lincoln Highway across the country is the important thing.’ Helen Joy’s monument . . . was surrounded by a square iron railing, anchored at each corner by a 1928 concrete marker post. Over the years, vandals shot or prized the bronze medallions off each one.”46 In 2001 the Lincoln Highway Association’s Wyoming members and the Wyoming Department of Transportation moved the monument to the Telephone Hill section of Interstate 80 east of Laramie, where sculptor Robert Russin’s enormous bronze head of Lincoln glares down at the superhighway that swallowed up his road.

Traversing the Desert

261

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 262

Interstate 80

Notes

Interstate 80, from New Jersey to San Francisco, completed in the 1970s, whisks travelers through the Red Desert at seventyfive miles an hour. There are 402.67 miles of I-80 in Wyoming, and it is heavy with trailer trucks hauling freight to the West Coast, machinery and pipeline to the gas fields of the west.47 Travelers through the Red Desert barely glance beyond the concrete roadway as the horizon stretches out and out and there is little to see beyond sagebrush and occasional pronghorns north or south of the highway’s long fence. The only surviving section of the old Lincoln Highway is U.S. 30 between Laramie and Medicine Bow, so hotly described by Elk Mountain residents as an inferior route. The engineers who laid out I-80 rejected the Medicine Bow route and ran their road just north of Elk Mountain, shorter in mileage and driving time. Elk Mountain is a formidable and huge presence, 11,156 feet in elevation, from whose heights sweep savage winds and winter storms. On I-80, shrieking winds tip semis over like toys and blinding whiteout snow often reduces visibility to a few inches beyond the windshield; black ice on the roadway is common. Colossal wrecks and multivehicle pileups occur every year in storms and bad weather. Travelers who know the local roads take the old Lincoln Highway Medicine Bow route when I-80 shuts down.48

1. Bullwhackers, like French-Canadian voyageurs, were singers. A favorite song was “Root Hog or Die,” which some scholars believe originated in New York State with ginseng diggers. 2. William Henry Jackson, Time Exposure (New York, 1940; Tucson, Ariz., 1994), 110 –111; page references are to the 1994 edition. Jackson’s notebooks were the basis for his autobiography. Year later he enlarged his sketches into paintings of life on the Oregon Trail, his sharp, observational eye invaluable to historians interested in the details of equipages, places, men, animals, and machinery. Much of what we know of this period in western history is built on his memories and paintings of ox teams, river towns, stage stations, springs, forts, stampeding bison, river crossings, miners, Indians, mountain men, cattle drives. He was there and he showed us how it was. Jackson is known primarily as the great photographer of the west from his work with the Hayden expedition of 1870 in photographing what later became the first national park in the world—Yellowstone. He spent nine years with the U.S. Geological Survey and was with Hayden’s party when it discovered the Mesa Verde cliff dwellings. For years Jackson ran successful photography studios in Denver and Omaha, eventually touring the world as the first international photojournalist. Much of his photography and many of his paintings recorded landscapes and life in early Wyoming. Most of the paintings are at the Scotts Bluff National Monument in Gering, Nebraska. 3. Ibid., 130. 4. Ibid., 102. The inmates of Wyoming Territorial Penitentiary in its thirty years of existence (1872 –1903) also show a rich distribution of national origins. See Elnora L. Frye, Atlas of Wyoming Outlaws at the Territorial Penitentiary (Jelm Mountain Publications, Laramie, Wyo., 1990). In addition to men and women from thirty-seven states and Indian Territory there were prisoners from England, Canada, Ireland, Germany, Scotland, France, Greece, Switzerland, Mexico, Italy, Austria, Sweden, Prussia, China, Norway, Spain, Denmark, Australia, Wales, Finland, the West Indies, Russia, Poland, and India. 5. Both Quien Hornet and Bishop mountains also appear on Clarence King’s 1876 map of the Green River Basin, in Engineer Department U.S.A. Geological and Topographical Atlas accompanying the report of the Geological Exploration of the Fortieth Parallel made by authority of the Honorable Secretary of War under the direction of Brig. And Brvt. Major General A. A. Humphreys Chief of Engineers U.S.A. by Clarence King U.S. Geologist in charge (Julius Bien Lith., 1876) and on many Wyoming maps after Stans-

262

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 263

bury. Quien Hornet also shows up on a list of mountains in Annals of Wyoming as recently as 1928 (vol. 5, 89), though placed in Uintah County. Bishop Mountain is likely today’s Pine Mountain, the highest point in Sweetwater County at 9,550 feet. The USGS Geographic Names office thinks Quien Hornet is probably today’s Little Mountain. The original name-giver of “Quien Hornet” is not known. Colorado scholar Francis Whittemore Cragin noticed the odd name on Stansbury’s map. His notes read: “‘Quien Hornet’ What would it mean? Do New Mexicans use the word hornet?”; Francis Whittemore Cragin Collection, Appendix A, Early local history, Series 2, Box 7, Colorado Springs Pioneers Museum. Although we do not know who named the mountain “Quien Hornet” (“Who is the hornet?”), the name seems connected to a traditional Mexican children’s game, Doña Blanca. In this game any number of children pretending to be pillars stand in a closed circle, holding hands. Inside the circle stands Doña Blanca, and outside is the child chosen as the “jicotillo,” or hornet. The pillars circle Doña Blanca, who skips about while the hornet ranges outside, looking for an opening. The pillars swing their arms and sing: “Doña Blanca está cubierta con pilares de oro y plata. / Romperemos un pilar para ver a Doña Blanca.” (“Mrs. White is surrounded by pillars of gold and silver. / Break open a column if you want to catch her.”) The next line is sung by Doña Blanca: “¿Quién es ese jicotillo que anda en pos de Doña Blanca?” (“Who is this hornet who chases Mrs. White?”) And the hornet replies: “¡Yo soy ese jicotillo que anda en pos de Doña Blanca!” (“I am the hornet who is trying to catch her!”) The hornet, unable to break into the circle, now asks various pillars of what material they are made. The individual pillars answer that they are made of gold, silver, and other unyielding substances until one pillar responds by claiming to be made of a softer substance, as wood, straw, paper, or cloth. At this point all the pillars drop their hands, and the hornet rushes in. It seems possible that a Spanish-speaking someone with a sense of humor named this mountain in the Red Desert after a “pillar” or Doña Blanca herself. The absence of the name on maps previous to Stansbury’s suggests that Bridger gave Stansbury the name. See Carl I. Wheat, Mapping the Transmississippi West (Institute of Historical Cartography, San Francisco, 1959), 3:124. 6. Wheat, Mapping the Transmississippi West, 3:125. 7. See Scott Thybony, Robert G. Rosenberg, and Elizabeth Mullett Rosenberg, The Medicine Bows: Wyoming’s Mountain Country (Caxton Printers, Caldwell, Idaho, 1986), 30 –31, 37–39. 8. A. Dudley Gardner, “Historical Overview and Evaluation of the Cherokee Trail, T12N, R97W, Sections 14 and 15” (Archeological Serv-

ices of Western Wyoming College, Rock Springs, 1981). 9. A. Dudley Gardner, Debbie Allen, and David Johnson, “Historical Assessment of the Cherokee Trail and the Bryan to Brown’s Park Road within the Wold Trona Lease Area” (Archeological Services of Western Wyoming College, Rock Springs, 1994). There have been subsequent and ongoing revisions of this assessment. 10. Patricia K. A. Fletcher, Jack Earl Fletcher, and Lee Whiteley, Cherokee Trail Diaries, vol. 1, 1849, A New Route to the California Gold Fields, and vol. 2, 1850, Another Route to the California Gold Fields (Caxton Printers, Caldwell, Idaho, 1999). In 2001 the Fletchers published an additional set of diaries: Jack Earl Fletcher and Patricia K. A. Fletcher, Cherokee Trail Diaries, vol. 3, Emigrants, Gold Seekers, Cattle Drives, and Outlaws (Thomson-Shore, Dexter, Mich., 2001). Lee Whiteley produced his own version; Lee Whiteley, The Cherokee Trail: Bent’s Old Fort to Fort Bridger (Johnson Printing, Boulder, Colo., 1999). 11. A. Dudley Gardner, Debbie Allen, and David Johnson, “Historical Assessment of the Cherokee Trail and the Bryan to Brown’s Park Road, in Sweetwater County, Wyoming” (Western Archeological Services, Rock Springs, Wyo., 2006); and David E. Johnson, “Historical Assessment of Cherokee Trail Segments along Cherokee Creek, Wild Horse Basin and Peach Orchard Flat, Carbon County, Wyoming” (Western Archeological Services, Rock Springs, Wyo., 2006). 12. Fletcher et al., Cherokee Trail Diaries, 1:124. 13. Today we suspect this illness may have been caused by ticks, including the Rocky Mountain spotted tick, or by Giardia lamblia, a singlecelled protozoan found in contaminated water used by bears and beaver and likely transmitted to humans who drink the unboiled water. 14. The Expeditions of John Charles Frémont, ed. Donald Jackson and Mary Lee Spence, vol. 1, Travels from 1838 to 1844 (University of Illinois Press, 1970), 214. On Fremont’s expedition to South Pass in 1842 –1843, the party temporarily split near the confluence of the Laramie and North Platte rivers, the site of Fort Laramie. The party’s cartographer, Charles Preuss, kept a journal in Fremont’s absence and noted the arrival at their camp on July 8 of Jim Bridger, “a man well known in the history of this country” (Preuss, Exploring with Frémont: The Private Diaries of Charles Preuss, Cartographer for John C. Frémont on His First, Second, and Fourth Expeditions to the Far West, trans. and ed. Erwin G. Gudde and Elisabeth K. Gudde [University of Oklahoma Press, Norman, 1958], 25). Bridger brought the unwelcome news that the Sioux, Cheyennes, and Gros Ventre had banded together in a war party scouring the country between Red Buttes and the Sweetwater, directly on the expedition’s planned route.

Traversing the Desert

263

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 264

Bridger himself, seeking to avoid the Indians, had not made his way from his trading post to Fort Laramie by the usual route, but “availing himself of his intimate knowledge of the country, he had reached [Fort] Laramie by an unusual route through the Black Hills, and avoided coming into contact with any of the scattered parties.” Jackson and Spence note that Dale Morgan suggested that Fremont apparently referred to “this entire area from Fort Laramie south to the Cache la Poudre as comprising a general range of ‘Black Hills.’” Daniel L. Kinnaman, in his Little Piece of Wyoming (Kinnaman Publications, Rawlins, Wyo., 1997), notes that he believes Bridger’s “unusual route had to be the Bridger’s Pass route.” If so, Preuss’s note would be the earliest reference to that pass, predating Stansbury’s by seven years. 15. Some of them insisted on following Fremont’s route north to the Sweetwater River and the California-Oregon-Mormon Trail, but rejoined Evans two days later. Fremont himself had had difficulty with the route, which took him over great and burning alkali flats. 16. Pyeatt, in Fletcher, Cherokee Trail Diaries, 1:119. 17. Brigham D. Madsen, ed., Exploring the Great Salt Lake: The Stansbury Expedition of 1849 – 50 (University of Utah Press, Salt Lake City, 1989), 651. 18. Crawford, in Fletcher, Cherokee Trail Diaries, 1:121. “Calebrate” was Crawford’s word for “Chalybeate,” meaning “water rich in iron.” 19. Ibid., 123. Dudley Gardner remarks that Calvin Holmes, who took the Cherokee Trail over Powder Rim in 1854, wrote that he rejoined the main stem of the California-Oregon-Mormon Trail along Blacks Fork about fifty-two miles east of Fort Bridger. Calvin Holmes, Diary, June 22 to July 10, 1854, Society for California Pioneers, Alice Phelan Sullivan Library, San Francisco. 20. Merrill J. Mattes, Platte River Road Narratives: A Descriptive Bibliography of Travel over the Central Overland Route to Oregon, California, Utah, Colorado, Montana and Other Western States and Territories, 1812 – 1866 (University of Illinois Press, Urbana and Chicago, 1988), 2. 21. Fletcher, Cherokee Trail Diaries, 1:94–95. 22. Ibid., 1:95–96. 23. Ibid., 1:96, 191ff. The Fletchers described the 1850 routes through the use of discovered diaries kept by William Quesenbury, James Mitchell, and John Lowery Brown. 24. Ibid., 1:243; “Jones came by us in the bottom with four squirrels— two of which were black.” 25. Ibid., 1:93. Kinney’s route was the same as Dick Owens’s from Pueblo to Brown’s Hole, though Kinney did not take the Ithaca party on

264

to Fort Bridger but went instead through the Uinta Mountains to Fort Uinta. Kinney hired out to the Ithaca party for $700 and his choice of two wagons and three mules. 26. Ibid., 1:330. 27. Brevet Brigadier General USV. The “USV” (United States Volunteer) may indicate that the general was a “galvanized Yankee,” a rebel who had changed sides. The inspectors included William Tecumseh Sherman, Orville E. Babcock, James Rusling, William B. Hazen, Rufus Ingalls, and Delos B. Sackett. 28. James F. Rusling, Across America; or, The Great West and the Pacific Coast (Sheldon and Co., New York, 1875), 143. 29. Ibid., 149. 30. Ibid. The alkali seems to have especially bothered the general, for he devoted a long paragraph to it: “Here, when the sun got fairly up, the sharp keen winds of the night hours changed to hot sirocco breezes, that laden with the alkali dust there became absolutely stifling. Alkali or soda—the basis of common soap—abounds throughout all this region for two or three hundred miles, and literally curses all nature everywhere. It destroys all vegetation, except sage-brush and grease-wood, and exterminates all animals, except cayotes and Indians. The Indians even mostly desert the country, and how the cayotes manage to ‘get on’ is a wonder and astonishment. The wheels of our coach whirled the alkali into our faces by day and by night, in a fine impalpable dust, that penetrated everywhere— eyes, ears, nose, mouth—and made all efforts at personal cleanliness a dismal failure. The only results of our frequent ablutions were chapped hands and tender faces— our noses, indeed, quite peeling off. In many places the alkali effloresced from the soil, and at a little distance looked like hoar-frost. It polluted the streams, giving the water a dirty milky hue and disgusting taste, and in very dry seasons makes such streams rank poison to man or beast. The Plains of Sodom and Gomorrah . . . could not have been much worse than this Desert of the Mountains” (ibid., 149 –150). 31. Ibid., 150 –151. 32. Ibid., 151. The thoroughbrace was one of the many inventions of the New England carriage makers in the late eighteenth century as they developed a rugged, flexible coach. Where English coaches, which ran on smooth roads, used steel springs, on New England’s stony roads such coaches capsized on corners and bruised horses’ shoulders and passengers’ rumps. Ralph Moody, in Stagecoach West (Thomas Y. Crowell, New York, 1967), 12, describes the innovation: “A heavy undercarriage was mounted on stout axles. . . . At the four corners of this frame iron stan-

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 265

dards about a foot high were mounted and securely braced. To the top of the pair on each side was shackled a six- or eight-ply belt of thick bull-hide leather about four inches wide, forming a cradle upon which the body of the coach was suspended.” 33. Robert Louis Stevenson, Across the Plains (Charles Scribner’s Sons, New York, 1897), 66 – 67. 34. Ibid., 46 – 47. 35. Ibid., 49. In San Francisco he was first diagnosed with tuberculosis, suffered a bout of malaria, and finally married Fanny Osbourne. Despite poor health he traveled the world, ending on the island of Samoa, where he lived before succumbing, at the age of forty-four, to a stroke caused by overwork and stress. His biographer, Frank McLynn, remarked, “All his utopias, whether Barbizon, Silverado or Samoa, ultimately disappointed him”; McLynn, Robert Louis Stevenson (Random House, New York, 1993), 178 –179. Nor did Charles Preuss, the brilliant cartographer with Fremont’s expeditions, find the Wind River Mountains impressive. For Preuss nothing in the west was as fine as his memories of Germany. In his diary entry for August 4, 1842, he wrote of the Wind River range: “Whoever has seen Switzerland and expects something similar here is bound for a great disappointment. An American has measured them to be as high as 25,000 feet. I’ll be hanged if they are half as high, yea, if they are 8,000 feet high. A little snow on their peaks; that is all, as far as I can see now, that distinguishes them from other high mountains” (Preuss, Exploring with Frémont, 33). 36. P. W. Spaulding, wire to A. R. Pardington, September 2, 1913, P. W. Spaulding Papers, Box 37, American Heritage Center, University of Wyoming (hereafter “Spaulding Papers”). There were subconsuls in the state as well. 37. The checkerboard railroad land distribution in the west caused many tedious right-angle detours or lease negotiations with the railroads. 38. Pardington to Spaulding, December 4, 1913, Spaulding Papers, Box 37. Bigger quarrels were to come, with San Francisco and Los Angeles fighting for the main route connections through Utah and Nevada. 39. P. J. Quealy had earlier told Spaulding, who passed the information along to Pardington in his long October 23, 1913, letter, “that he had three of the gates in use but if the road was selected he would take them out and build eight miles of fencing in their place and move his house if we desired etc.” Spaulding Papers, Box 37. 40. John Hay to Spaulding, November 21, 1914, Spaulding Papers, Box 37. 41. Pete Davies, American Road: The Story of an Epic Transcontinental

Journey at the Dawn of the Motor Age (Henry Holt and Co., New York, 2002), 141. 42. Ibid., 145. 43. Richard Reinhardt, Out West on the Overland Train (Castle Books, Secaucus, N.J., 1967), 79. 44. Gregory Franzwa, The Lincoln Highway, vol. 3, Wyoming (Patrice Press, Tucson, Ariz., 1999), 29. Franzwa guides the reader along the old 1913, 1921, 1924, 1935, 1940, and 1950s routes through the Red Desert section of Carbon County, based on the pieces of highway that lie beneath existing Route 30: Medicine Bow, Carbon, Hanna, Elmo and westward. 45. Davies, American Road, 222. 46. Ibid., 229 –230. A local resident who remembers the monument well remarked that an entrepreneurial vandal from Rock River prized the medallions off and mounted them on bookends, which he sold (pers. comm.). 47. Bob Moen, “State Has Plans for I-80,” Rawlins (Wyo.) Daily Times, April 20, 2006. 48. There is a local legend to the effect that while the highway was under construction Lady Bird Johnson, on a tour of the west, noticed the beautiful wildflowers on the north side of Elk Mountain and used her influence to request that the route run along the mountain’s base so that travelers could enjoy the blooms. Oddly enough, this writer heard a similar Lady-Bird-wild-flower-highway-route story in Oregon.

Traversing the Desert

265

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 266

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 267

16. Forts of the Red Desert Annie Proulx

ive of Wyoming’s sixteen forts and camps stamped their character on the Red Desert: Forts Bridger (1842/1857– 1890), Halleck (1862 –1866), and Fred Steele (1868 –1886) and Camps Stambaugh (1870 –1878) and Pilot Butte (1885– 1898). So-called “Fort” LaClede, although in the Red Desert, was never a military fort.1 The liveliest phase of military fort construction in the American west, from the 1840s into the 1880s, paralleled the country’s expansionist successes—the annexation of Texas, settlement of the Oregon question with England, and the incorporation of the rich territories gained through the Mexican-American War. This was the time of American empire building and government centralization, the time of exploration of the west’s resources, the search for passes through the Rockies, a major emigration surge, and several gold and silver rushes. Much of this action touched the Red Desert—strategically located so that railroad, emigrant, telegraph, sheep, and cow had to cross or inconveniently skirt the area. During the boom days of fort building, soldiers did the work. But as the reservations corralled the Indians, fewer troops were needed at fewer forts, especially when messages could be quickly sent by the “talking wire.” In the decade 1870 –1880, military troops represented 23 percent of the Wyoming labor force.2 After 1868 the railroad made it possible to rush troops close to trouble

F

spots, reducing the necessity for so many forts. As the reservation system entrenched itself, the army gave fort-building contracts to civilians instead of relying on troop labor. Bridger, Halleck, Steele, Pilot, and Stambaugh had very different characters and histories. All of them attracted and interacted with the civilian life that sprang up around them, and not a few soldiers, after they were mustered out (or deserted), themselves became ranchers, prospectors, cowboys, storekeepers, freighters, and saloonkeepers in the territory. During their enlistment periods they had learned the vagaries of the climate and the lay of the land. They had fought Indians, escorted freight wagons and emigrant trains, repaired miles of telegraph line, defended stage stations, tracked outlaws, guarded property from striking miners, executed criminals, hunted big game, fished pristine waters, and experienced a wild assortment of extreme weather. Some understood the opportunities. For the officers, a tour of duty at these forts was often one long hunting party. The bison became scarce in the 1870s and by 1880 were more or less gone, wiped out by hide and sport hunters and, possibly, diseases carried to them by the trail herds, but there were still huge numbers of antelope, elk, and deer, plentiful mountain sheep, immense flocks of ducks, and countless greater sage-grouse. For men to whom hunting was a primary pleasure, this unrestricted 267

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 268

Camps and forts in the Red Desert

world of game was addictively thrilling. Fort commanders often played host to visiting friends and political nabobs from back east. A fishing party on Battle Creek took 1,200 fat trout, with Fort Steele’s commander, Major Thornburgh, catching 56 in thirtytwo minutes.3 Life at the frontier forts was hard, rough, and dangerous. Everything was difficult. The men had to cut and transport timber, build their own shelter. Troops suffered from scurvy, pulmonary infections, dropsy, various fevers, wounds, poor medical facilities, the psychological stress of isolation, fear of Indian attack, the ceaseless wind, and horrible weather. The order book of Colonel William O. Collins, commander of Fort Laramie, for October 1862 –April 1863 is permeated with anxieties over shortages of hay, wood, corn, and ammunition, especially at the small 268

forts and posts under his command. Collins refers to a man in Company F who cracked and became “crazy and unmanageable.” The winters, more severe than those of today, killed many of the inadequately clothed men. Long, thirsty marches through arid high desert and battles with an enemy fighting for its homeland both encouraged and discouraged desertion. During the 1860s the Eleventh Ohio Volunteer Cavalry manned the forts of the Rocky Mountain frontier. The regiment never served together during the period but was scattered among the forts and posts along mail and emigration routes.4 They were involved in the 1862 South Pass battle, the Sweetwater Bridge fight in 1863, and in 1865, “the bloody year on the plains,” fought in almost two dozen Indian battles. The tribes they came up against practiced an extreme form of democracy to whom the idea

Annie Proulx

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 269

of disciplined companies, regiments, and military hierarchy was alien. Individual achievement and glory through a loose amalgamation of warriors better characterized their style of fighting. The first battalion, all Unionist Ohioans, arrived at Fort Laramie in 1862. The next year, the second battalion arrived, composed of companies E, F, G, and H, and among its men were a few “galvanized Yankees,” Southern rebels who had been captured and who, in exchange for release from prison, took an oath of allegiance to the United States and enlisted in the Union army. In June of 1864 the so-called surplus recruits assigned to the regiment were organized into companies I, K, and L. Many of these men were ex-POW, galvanized Yankees. Some of the men went by assumed names, some plotted mutiny and were discovered and shot, some deserted and made western lives for themselves, though most served well and honorably. These men contributed to the early character of Wyoming.

Notes 1. Louise Bruning Erb, Ann Bruning Brown, and Gilberta Bruning Hughes, The Bridger Pass Overland Trail 1862 –1869 through Colorado and Wyoming and Cross Roads at the Rawlins-Baggs Stage Road in Wyoming (erbgem, Littleton, Colo., 1989), 94. “Ft. LaClede had an official military designation as a fort. In 1863 it was garrisoned by four companies of the 11th Ohio Cav. Volunteers. Troops were on scout and escort duty at all times patrolling from Ft. LaClede east to Sulphur Springs and west to Green River. In 1865 these men were relieved by the 11th Kansas Cav. under the command of Col. P. B. Plumb and were involved in a great deal of Indian activity during the life of the trail” (ibid.). Robert W. Frazer does not list Fort LaClede in Wyoming in his Forts of the West (University of Oklahoma Press, 1965, 1972), nor do other historians of the military west. Corroboration for the Eleventh Ohio and Eleventh Kansas Cavalry occupation of this “fort” has not turned up in Fort Laramie archives. The Erb-Brown-Hughes assumption that the ruins were once a fort may be based on a 1981 report done for the Amoco Corporation by Robert G. Rosenberg and Peter Kvietok, “An Historical Overview of the Red Desert Region” (Archeological Services, Laramie, 1981), 11–12. In 1995 Russel Tanner‘s “Who Built the Little Rock Houses: An Ethnoarcheological Study of Overland Stage Stations in Southwestern Wyoming” (master’s thesis, University of Wyoming, Laramie), clarified the identities of the stage stations, particularly the Salt Wells station. James H. Nottage, in “Defending Western Trails, Arms of the Eleventh Ohio Volunteer Cavalry, 1862 –1865” (American Society of Arms Collectors, 1997), had not found any references that put the Ohio Volunteers at LaClede. 2. Lewis L. Gould, Wyoming: A Political History, 1868 –1896 (Yale University Press, New Haven, Conn., 1968), 15. This figure, says Gould, compares with a national average of two-tenths of 1 percent in the period. 3. Taylor Pennock, “Recollections of Taylor Pennock,” Annals of Wyoming 6, nos. 1–2 (July–October 1929): 208. John C. Van Dyke spent much time cowboying in Wyoming and Montana in the 1870s. His Open Spaces: Incidents of Nights and Days under the Blue Sky (1922; reprint, University of Utah Press, 1991), will make today’s sportsmen and game managers sick with the sense of loss. 4. Theoretically there were 100 men in a company, three companies plus support troops in a battalion, three battalions in a regiment, 5,000 men in a brigade, and 15,000 in a division. The frontier companies of the Eleventh Ohio rarely reached the full complement of 100 but wavered between 69 and 96 men in the 1860s. Although subdivisions of the company, platoons, were known before the Civil War, on the frontier the smallest unit was usually the company.

Forts of the Red Desert

269

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 270

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 271

17. Fort Bridger and Camps Stambaugh and Pilot Butte Dudley Gardner

ort Bridger was one of the longest continually occupied sites in southwest Wyoming. In its soils is the record of changes that Native Americans, Europeans, and Americans made to the region over two centuries. The newcomers sowed exotic plants in the fertile valley irrigated by the Blacks Fork River. Weeds and wheat grew where they had never grown before. Hay from these fields fed cattle and horses that were traded to trail-weary westering emigrants. The cattle industry had a toehold in the region by 1843. The fort, also a trading post, stood near shifting international borders on tribal lands claimed by the Bannocks, Utes, and Shoshonis. In 1843 the area was claimed by Mexico, just south of British territory and southwest of the United States. With the end of the Mexican-American War in 1848, the fort sat just inside the northwest corner of Utah Territory, not far from the territorial boundaries of Oregon and Nebraska. From 1848 to 1868 the fort was, by turns, inside Utah, Dakota, and finally Wyoming Territory. The various owners of the fort pursued diverse relations with the Native Americans. Jim Bridger saw them as friends or enemies, but always as trading partners. They served as a buffer between himself and the Utah settlement’s theocratic Mormon government. This was Bridger’s third fort with the advantage of broad pastures and easy access to the Oregon Trail. In 1847 Mor-

F

mon immigrants stopped at the fort, and during the 1849 gold rush, California-bound travelers traded for food, blankets, and moccasins at the post. The post owed much of its success to the Shoshoni and Ute wives of the traders who lived at the fort. They sewed moccasins, prepared food, and made leather goods for trade. The women were excellent traders, and some remained at the post even after the Mormon Church took over the fort in 1855. Jim Bridger himself was at one time married to a Shoshoni woman. From roughly 1843 to 1853, Bridger, or, more likely, his Native American partners, ran the trading post. Bridger was something of a passive owner. For example, he would at times lend his blacksmith tools and shop rather than working in the facility himself. We know this because in 1847 one of the emigrants using the shop burned the building down. This was hard on business but good for the archaeological record, as the ash from this fire covers portions of the floor. Following the 1847 fire, the residents of the post put a sand cap over the ash. Both the ash and sand are clearly identifiable stratigraphic markers of the past. Joel Palmer, traveling west in 1845, camped near the trading post on July 25. He described the post as “built of poles and daubed with mud; it is a shabby concern.” He went on to say that “there are about twenty-five lodges of Indians, or rather white 271

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 272

trappers lodges occupied by their Indian wives.” They had “dressed deer, elk, and antelope skins, coats, pants, moccasins, and other Indian fixens, which they trade low for flour, pork, powder, lead, blankets, butcher-knives, spirits, hats, ready made clothes, coffee, sugar, and etc.” The traders’ wives were, according to Palmer, “mostly of the Pyentes and Snake Indians.” For a horse, the going rate at Fort Bridger was “twenty-five to fifty dollars in trade.” Since horses were often worn down by the time they reached the Bridger Valley, there was a substantial market for fresh horses.1 In 1853 certain events forever altered the history of the post. Brigham Young, the U.S. Indian agent for the region, issued orders to have Jim Bridger arrested for selling alcohol to Native Americans. Bridger fled the Mormon posse. What happened next is cloudy, but by 1855 the Mormon Church had purchased Bridger’s trading post from Louis Vasquez, Bridger’s partner. They began making alterations to the structure in 1857. The old picket walls were removed, Bridger’s home remodeled, and a stone fortress erected around the compound. The Mormons continued to operate the trading post and actively courted Shoshoni patrons. This all came to an abrupt end in 1857. Fred R. Gowans, Eugene E. Campbell, and Dr. Norman F. Furniss pointed out the Mormons’ policy toward the Indians in Utah Territory. First, and most important, since the Mormons viewed the world through eyes focused on their faith, they had an obligation to convert the Native Americans to their religion. They wanted to maintain positive trade relations with the Indians; build settlements on land claimed by the Shoshonis, Utes, Gosiutes, Paiutes, and Bannocks; and use the Indians as buffers between them and the various outside groups pressing against the borders of the State of Deseret, or Utah, as the federal government officially designated the territory. Finally, they wanted to ensure the peaceful passage of Mormon emigrants from the Missouri River to the Salt Lake Valley. Fort Bridger was the key. In 1857 the so-called Mormon War ended with the Latter-day Saints burning Fort Bridger to the ground and the U.S. Army seiz272

ing the post.2 Like the 1847 fire, the 1857 fire created a unique archaeological signature; moreover, the army leveled the interior by capping the burned deposits, an effective seal that covered most of the interior of the compound. The army also revamped the stone fort and built bastions to enhance the defensive capabilities of the structure. These bastions and stone walls stood in place until the 1880s, when the entire structure was razed, and enlisted men’s barracks constructed atop the old compound. All of the fires, remodeling activities, and new construction were outlined by clear-cut stratigraphic markers.3 My colleagues and I analyzed the type and variety of lithic artifacts we found in the two occupations dating from 1843 to 1855 when Bridger owned the post, and during the Mormon occupation from 1855 to 1857. Throughout everything, Shoshoni Indians lived and worked at the fort. While all excavated levels contained lithic materials in the fill, it was in the Bridger and Mormon occupation zones that the greatest number and variety of stone tools were found. In all, nine stone (lithic) projectile points were recovered. In the Bridger occupation level, five projectile points — classified as Rose Springs/East Gate, or Uinta phase—were uncovered in direct association with gun flints, pins, rifle parts, percussion caps, and a variety of items dating to the mid-eighteenth century. Also two desert side-notched, one cottonwood series point, and a projectile point similar to an Elko point came from the excavated area.4 The Rose Springs, or corner-notched, projectile points came from a stratigraphic level that dates from 1857 to 1890. There is a high degree of reliability that all five points were deposited from 1843 to 1857. The desert side-notched points may have been deposited somewhat later, from 1857 to 1890. It appears that corner-notched points enjoyed a longer period of use than previously thought. In fact, there seems to be a stylistic continuity in corner-notched projectile points that spans about 1,100 years. The Mormons who gained control of the fort in 1855 truly saw the Shoshonis as trading partners and catered to their needs, which included trade beads. Later they traded barley, wheat, and potatoes to several tribes.

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 273

Numerous travelers bound for Oregon, Utah, and California described the trading post and even made sketches of it. Unfortunately, the sketches and descriptions confuse rather than clarify the nature of the trading post Bridger and Vasquez constructed. The numerous diaries and sketches, when combined with the results of archaeological excavation, provide a great deal of information about the lives of the post’s inhabitants. It appears the post was constructed with two enclosures: one for post inhabitants and one for livestock. This configuration was modified when the fort was purchased by the Mormon Church in 1855.5 Lewis Robison, laboring on behalf of the Church, constructed a stone fort from river cobbles. Because the Mormons built their compound on top of the original trading post, defining the first walls of Bridger’s post is somewhat difficult. It does appear that the Mormons reused the structures built by Bridger and Vasquez, alluded to in Bridger’s “Spoliation to the United States Government,” a claim for damages filed years after the destruction of the Mormon compound. The archaeological record supports this construction. Bridger’s original buildings had earthen floors, which are archaeologically discernable as a well-compacted occupation floor. Atop this floor are artifacts congruent with the 1840s, but more important, the original floor area is sealed or covered in places with a thin layer of ash in the southwest corner of the compound. In 1847 William Clayton, who chronicled the initial Mormon migration west to the Salt Lake Valley, reported that Bridger’s blacksmith shop accidentally burned due to a careless visitor who had misused his forge. The Mormon remodeling laid down wooden planks for flooring. In 1857, when the Mormon compound burned, the flooring also burned. The military who occupied the post in the same year simply leveled the burned rubble and constructed their buildings on top of it. The two burning episodes provide distinct time markers that allow us to separate the different occupation levels. During excavation, pollen samples were taken from above and below the

burned planking. The pollen analysis shows numerous changes had been initiated by both the traders (1843–1855) and the Mormons (1855–1857). Atop the wooden floor both barley and wheat pollen were recovered.6 The historical record indicates that the Mormons successfully raised wheat and potatoes in the valley. 7 This valley is located at roughly 6,300 feet above sea level and has a relatively short growing season. The Mormons’ successful cultivation of wheat near Fort Bridger is significant. It provided the fort’s inhabitants with food and even a surplus they could trade to the Shoshonis in the area. Two manos were recovered in 1991. Both appear to be directly associated with the occupations spanning 1843–1857. The manos indicate the use of stone tools to process grain in the fort, and to date the only indication of grain within the compound is domestic wheat and barley. Cultivation of such exotic plants marked the beginning of a permanent alteration in the native plant communities of the Blacks Fork Valley. Linda Scott, who analyzed the pollen at Fort Bridger, notes that during the Bridger and Mormon occupations there is a relatively small amount of weed pollen in the area sampled, but “increases in pollen that might represent weed[s] and [a] decline in sedge and grass pollen during the military occupation suggest the military personnel used this area more heavily than did the civilians.” This heavier use created “greater ground disturbance and possibly” led to “wearing bare areas that were once grassy areas,” thus contributing to weed growth.8 Within the Mormon and Bridger occupation levels, recovery of charred Prunus (cherry) and Sorbus (mountain ash) seeds indicates that the occupants had access to cherries, “possibly as canned cherries, and suggests the possibility that a shrub such as mountain ash grew in the area. It is most likely that these seeds were accidentally charred when the Mormons burned the compound prior to leaving.”9 Among the other nonarboreal pollens, Scott noted sunflower, dandelion, chicory, Mormon tea, wild buckwheat, strawberry, and rosehips (Rosaceae).10 The 1992 excavations identified what appears to be a dug pit made during Bridger’s occupation of the fort. Within this pit a

Fort Bridger and Camps Stambaugh and Pilot Butte

273

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 274

gun flint, parts of an iron kettle, and the remains of a schist blank (used to make ornaments or implements) were found.11 All of these were valued trade items. Corn and biscuit root starch also were found within this storage feature.12 This represents a blending of diverse cultural foods, one exotic, the other growing in the Wyoming desert. Biscuit root was a valuable food source for Native Americans living throughout the region. Corn had probably not been cultivated in the Bridger Valley prior to the arrival of Euro-Americans in the 1840s. Its presence in the storage feature presents interesting questions such as whether corn was grown locally or imported from Taos, the Missouri River valley, or the Salt Lake Valley. Whatever the answer, the presence of this exotic has intriguing implications. The “Mormon War,” or “Utah War,” that ended the Mormon occupation of Fort Bridger reflects the politics of the midnineteenth century. The actual war occurred only in southwestern Wyoming and no one died in the conflict, but it changed the region. In 1857 President James Buchanan ordered Brigham Young to step down as territorial governor of Utah. Young, of course, also headed the Mormon Church. To enforce his edict, Buchanan ordered troops to accompany the newly appointed governor, Alfred Cumming, to Utah. This sent terror into the young Mormon colony in Utah. In 1847 they had fled west to avoid attacks like the one at Nauvoo, Illinois, which ended in the death of their prophet Joseph Smith. Since their arrival in Salt Lake Valley in 1847, they had been left pretty much alone. That changed in 1857.13 General Albert Sydney Johnston wrote that on November 6, 1857, winter struck en route to Fort Bridger. The cold racked “the bones of our men,” he related, adding that “our oxen, and mules and horses, already half starved . . . died on the way and at our camps by the hundreds.”14 He noted that the thermometer hit sixteen degrees below zero. They reached Fort Bridger by November 17.15 The loss of animals led him to dispatch “Captain Randolph B. Marcy to Fort Union, New Mexico for droughtmules, and a remount for dragoons and batteries.”16 Marcy returned to Fort Bridger by way of Bridger Pass. The success of this 274

expedition through the pass led, in the words of General Fitz-John Porter, “to the opening of the route . . . through Bridger’s Pass and down Bitter Creek” by the Sixth Infantry. In the future the military would use this route extensively, as the road was, in Porter’s opinion, “shorter, easier, and better for grass.”17 Brigham Young became convinced that the best way to defeat the army was to starve it out. Winter arrived in a fury, with temperatures dipping to fifty below zero. He reasoned that with no food and no fodder, the army would starve. From South Pass to the west, Young ordered the grasses and the supply wagons burned. To his horror, the military kept coming. He then ordered the burning of Fort Bridger, Fort Supply, Supply City, and all Mormon housing and settlements in the Blacks Fork Valley.18 Only a few months after the stone walls at Bridger had been completed, the post was burned. The winter of 1857 and 1858 led to a negotiated peace, with Cumming installed as governor of Utah.19 The military remodeled the post, which was to become a major military fort with improved supply roads. They put a great deal of effort into improving the roads to the fort. With the onset of the Civil War in 1861 and the switch of the Overland Mail south from the South Pass route to the Overland Trail, patrols from Fort Bridger kept the road from Bridger Pass to the fort open. Fort Bridger’s foremost role of the 1860s was to ensure the security of transports and emigrant trains. Native Peoples continued to trade there. In 1868 the Treaty of Fort Bridger placed the Shoshonis on the Wind River Reservation in central Wyoming. On paper it ended the Shoshoni claim to southwestern Wyoming; in reality the Shoshonis continued to roam freely throughout the region well into the twentieth century. Fort Bridger served as a place where the Shoshonis came to camp and trade long after the 1868 treaty. They camped near the old Mormon compound and traded with emigrants and settlers in the area. The archaeological evidence suggests that trade continued into the 1880s. In 1890, when Fort Bridger was abandoned as a military post, the Shoshonis seemingly stopped visiting the area to trade.

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 275

At the onset of the Civil War, Fort Bridger troops moved east, leaving only a few men to guard the fort. The situation at Fort Bridger in 1861 became complicated by the lack of troops. Judge William A. Carter, probate judge, camp sutler, and veteran of the Seminole War in Florida, raised a volunteer company. Numbering forty to sixty men, they guarded the post and the Overland Trail.20 Not until December 1862 did relief arrive when Captain M. G. Lewis with Company I, Third California Infantry Volunteers, assumed command of the post. For the duration of the war the fort was garrisoned by California and Nevada volunteers.21 Fort Bridger, between 1867 and 1868, also provided troops to protect the construction and survey crews along the Union Pacific Railroad. Lieutenant Colonel H. A. Morrow and the 36th Infantry at the post had responsibility for protecting South Pass and the Overland Mail as well as the railroad, an area extending two hundred miles east of Green River north to the Wind River Mountains.22 Military Camps Military camps were constructed to serve hot spots in the area. Camp Stambaugh (1870) and Camp Pilot Butte (1885) were both constructed in response to problems. At South Pass, miners in the 1860s had settled on land claimed by the Shoshonis. The claim was contested by the Sioux, Arapahos, and Cheyennes, and it would be these three groups that attacked miners in the South Pass area on a regular basis. Camp Stambaugh came into being to protect the gold miners. Camp Pilot Butte, built in Rock Springs, was different. The need to protect the Union Pacific Coal Company’s interests and the Chinese coal miners led to the creation of an urban military camp, a rarity in the west. Expansion of prospecting activity at South Pass led to increased hostility with the Arapahos who were in the region in the late 1860s. Along with Sioux and Cheyenne war parties, they were a direct threat to the three mining communities—Atlantic City, Miners Delight, and South Pass City. In 1867 a war party of 150

Sioux and Cheyennes rode through the area. In April 1870, three hundred “hostile Indians” fought a battle with miners thirty miles below present-day Lander.23 To calm the miners’ fears, regular army units from Fort Fred Steele and Fort Bridger were assigned to patrol the area. In May of 1870 Lieutenant Charles Stambaugh led a detachment patrol from a camp near Atlantic Gulch. They pursued a small band of warriors and ran straight into an ambush. Stambaugh died in the battle. A few weeks later the army established Camp Stambaugh east of Atlantic City in Smith Gulch.24 Named after the lieutenant who had died a month earlier, the post was under the command of Fort Bridger.25 Except in summer months, duty at the fort proved to be an exercise in keeping warm. The post consisted of two dozen log and wood buildings.26 John McDermott, a military historian, notes that “the two barracks were L-shaped buildings of shingled logs, set on log ‘sleeper’ foundations.” The barracks each had a “brick fireplace at one end, a store room, an orderly room lavatory and a kitchen.” For a mere camp, Stambaugh had substantial buildings with officers’ housing, a bakery, storehouses, a hospital, a guardhouse, a quartermaster building, a commissary, a grain house, an adjutant’s office, a carpenter shop, a stable, a slaughterhouse, and a blacksmith shop. Springs and two wells ensured good sources of water.27 Winter brought boredom. In a place where winter lasts from September to May, extreme cold and wind kept the men indoors. At 7,700 feet above sea level, Stambaugh was subject to winds whipping through it. The first year, twenty-foot snowdrifts on the parade ground forced the men to tunnel from barracks to other buildings. In fact, freezing to death became a real potential hazard of duty at Camp Stambaugh.28 In summer the routine changed. Shoshonis allied with the troops in an attempt to drive Sioux, Cheyennes, and Arapahos from lands the Shoshonis traditionally claimed.29 The military aided the Shoshonis, who carried the attack to the Cheyennes and Arapahos.30 Attacks increased in 1870. In the summer of 1870, “Indians killed Oliver Lamoureaux on the Point of Rocks Road

Fort Bridger and Camps Stambaugh and Pilot Butte

275

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 276

some 30 miles south of Atlantic City.” Not all attacks were on miners or freighters. On October 30, 1876, a village of five hundred Shoshonis “was attacked by a large Sioux war party” ninety miles from Camp Stambaugh at Point of Rocks. The end of the “Sioux War” in 1877 spelled the end of Camp Stambaugh, and in 1878 the army abandoned the camp. In May 1881 the army turned the post and the unofficial square-mile reservation it controlled over to the Department of the Interior. The Chinese Massacre and Camp Pilot Butte The origins of Camp Pilot Butte are distinct. Built in 1885, Camp Pilot Butte was situated next to the Rock Springs Chinatown, between what are today Elias and Bridger streets, to quell racial tensions between the Chinese and other nationalities in town. The need for this camp grew out of long-standing labor strife centered on the Union Pacific Railroad Company’s coal mines in Rock Springs. In 1884 the newly appointed president of the Union Pacific Railroad, Charles F. Adams Jr., became aware of labor problems in the west. Some observers saw Adams, a member of the distinguished Adams family that had produced two American presidents, as a possible warrior for the common man. A man in Cheyenne wrote to Adams asking for relief from Union Pacific’s reduction in wages, stating, “I am a member of the Knights of Labor, an organization which comprises many many thousands in the United States.” Claiming that the Adams name was “identical with Patriotism and the Peoples’ rights,” the writer asked for help in solving labor problems in Wyoming.31 Letters to Charles Adams came not only from members of the Knights of Labor but also from Union Pacific managers concerned with the running of the mines. S. R. Callaway, the general manager for Union Pacific Railroad at Omaha, became concerned about the growing anti-Chinese sentiment and pro-labor stance of the workers in Wyoming. He wrote Adams in January 1885: “[At] Carbon, the men were all out 276

on strike because we will not discharge the foreman and every one else who does not suit them. They claim that before they will go to work they are going to compel us to discharge all Finlanders and Chinese.”32 In Wyoming, “Finlanders” were considered “Mongolians,” or at least not of European ancestry. Callaway went on to say that if the “Knights of Labor order a general strike, I confess I do not see my way out.” Yet he also clearly pointed out, “I have told the men that we will not discharge the Chinese and have ordered the discharge of the men at North Platte and those engaged in the trouble last night.”33 Although the trouble was never clearly specified, the company’s policy was that the Chinese would stay. The Knights of Labor countered, “The Chinese must go.”34 Faced with two opposing forces, the Chinese found themselves in the middle of a brewing storm. The situation was obviously deteriorating. By the summer of 1885 the problem reached a crisis stage. John L. Lewis (not related to the labor leader), writing on behalf of the miners, notified D. O. Clark, superintendent of the Union Pacific Coal Department, “Although I have been lying sick in my bed for the last four weeks, I have been flooded with correspondence from Wyoming . . . the sum and substance of which is, that the Chinese are having all the work they can do, working night and day, whilst our men at Rock Springs are left out in the cold.” Clark added, “I understand that they are now working almost day and night, whilst Carbon men have worked but one day in the last two weeks. This makes the situation terribly aggravating, and in spite of my efforts will undoubtedly result in a severe struggle if long continued.” Then, in pleading terms, Clark asked: “For God’s sake do what you can to avoid the calamity; the pressure is more than I can bear. See that justice is done to all the men at Carbon, and to the unemployed portion at Rock Springs.”35 Union Pacific later held that Lewis’s letter did not arrive until too late. But the Union Pacific officially knew that labor problems were brewing along the Wyoming section of the main line. In early September the situation in Rock Springs worsened, and on September 2, 1885, in the tiny coal camp of Rock Springs, twenty-eight Chinese were murdered. The riot began in a room

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 277

in Union Pacific No. 6 mine. By tradition, once a miner began working a room it was his and off limits to other miners. On September 1, two miners, known in the historical record as “Whitehouse and Jenkins,” were assigned rooms in Entry No. 5. The pit boss gave them the first marked-off rooms without being clear as to where he wanted them to begin working. As the miners came to the first two rooms where working had not commenced, they began preparing the room for extracting coal. The next day when Whitehouse returned to work, he found “two Chinamen in possession of what he considered his room,” and feeling it was his room, he ordered them out. The Chinese miners, who were equally possessive over what they considered had been their assigned room, would not leave. According to the Rock Springs Independent, “High words followed, then blows. The Chinese from other rooms came rushing in, as did the whites, and a fight ensued with picks, shovels, drills, and the needles for weapons.” The newspaper, which stood behind the white miners, claimed the “Chinamen were worsted, four of them being badly wounded, one of whom has since died.”36 The Chinese workers in Union Pacific No. 6 mine differed from workers of other nationalities in that their ultimate goal was not to stay in America but to return to China. The Chinese held that “falling leaves from trees settle on their roots,” and like those leaves, people residing elsewhere finally must return to their ancestral home. Most Chinese made provisions, once they entered America, to have their bones shipped home if they lost their lives abroad. Tied to different cultural traditions, Chinese working in Rock Springs did not understand the nature of the other miners’ problems. They were not asked about their feelings, nor did they articulate them in the local press, as the white miners had.37 Moreover, there was a distinct language barrier, with few Chinese speaking fluent English. To the Chinese, the worst situation they could face would be the lack of work. For them, a strike meant time without pay and a delay in returning home to China. It was not that they were uncaring about conditions in the mines; it was, instead, a deep-seated concern that they might lose their jobs that

prevented the Chinese from joining the white miners’ labor union. The Chinese left No. 6 mine after the underground incident and returned to their company-owned houses on the north side of Bitter Creek. The foreman at No. 6 had ordered work to cease after the riot broke out underground. The white men apparently took refuge in the bars, and the Chinese in their homes.38 The Chinese later told officials, “About two o’clock in the afternoon a mob, divided into two gangs, came toward Chinatown.” One “gang” crossed Bitter Creek on a plank bridge, the other crossed over the railroad bridge, and they converged on the Chinese. Part of the mob moved to No. 3 mine, which was north of Chinatown. This meant that the Chinese were virtually encircled. Across Bitter Creek and over the ridge north of town were the only directions they could escape. As one report put it: “At that time the Chinese began to realize that the mob was bent on killing. The Chinese, though greatly alarmed, did not yet begin to flee.”39 The Chinese account of what took place next depicts a scene of mob violence. Some who were able to run away dispersed on the south side of Bitter Creek. Others fled to the north into the desert hills behind Chinatown. A number were unable to escape. One newspaperman reported: “Between 4 o’clock and a little past 9 o’clock p.m. all the camp houses belonging to the coal company and the Chinese huts [were] burned down completely.” The Chinese who witnessed the event reported: “Some of the Chinese were killed at the bank of Bitter Creek, some near the railroad bridge, and some in ‘Chinatown.’ After having been killed, the dead bodies of some were carried to the burning buildings and thrown into the flames.” The buildings that burned had dirt roofs, and these collapsed, killing those who tried to hide.40 Eyewitnesses told investigators that “some of the Chinese who had hid themselves in the houses were killed and their bodies burned; some, who on account of sickness could not run, were burned alive in the houses.” Another Chinese died “in ‘Whitemen’s Town’ in a laundry house, and his house [was] demolished.” The Chinese witnesses—who gave the most accurate accounting of dead

Fort Bridger and Camps Stambaugh and Pilot Butte

277

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 278

and injured— contended: “The whole number of Chinese killed was twenty-eight and those wounded fifteen.”41 Another report estimated that one hundred houses erupted in flames on September 2, 1885.42 Most of the surviving Chinese fled westward toward the town of Green River. Taking quick action, Union Pacific told the railroad conductors to pick up any Chinese they found along the railroad line. All the Chinese who reached the main line were given passage to Evanston, where it was said they could be protected. This was a questionable notion since the white miners at Almy, north of Evanston, were unsympathetic to the Chinese. The Salt Lake City Tribune reported: “The native citizens there threatened day and night to burn and kill the Chinese. Fortunately, United States troops had been ordered to come and protect them, and quiet was restored.”43 Company A of the Ninth Infantry and Company I from the Twenty-first Infantry, along with a Gatling gun, were then stationed in Evanston. Later, the Gatling gun and Company I would be moved to Rock Springs.44 One week later the troops escorted the Chinese from Evanston back to Rock Springs. It fell to Chinese immigrants to rebuild their community and reopen the mines. In 1880 there were 763 residents in Rock Springs, 497 of whom were Chinese.45 The Chinese majority steadily increased until 1885. Therefore, when Chinatown was burned in the riot, the fire destroyed much of Rock Springs. When the Chinese fled on September 2, not enough coal miners remained in town to operate the mines. The Union Pacific Coal Company desperately needed the Chinese to return to work to provide fuel for its trains. The military promised the company protection. According to a government report, when the Chinese arrived on September 9, they “saw only a burned tract of ground to mark the site of their former habitations.” Returning survivors found some dead still scattered in the ashes of Chinatown. The Chinese survivors reported: “By this time most of the Chinese have abandoned the desire of resuming their mining work, but inasmuch as the riot has left them each with only the one or two torn articles of clothing they have on their persons, and as they 278

have not a single cent in their pockets, it is a difficult matter for them to make any change in their location.” At this point the Union Pacific Coal Company “promised to lend them clothing and provisions, and a number of wagons to sleep in.” Even though the army protected the Chinese, “their sleep [was] disturbed by frightful dreams, and they [could not] obtain peaceful rest.”46 The Chinese returned to work only under the protection of federal troops. The Chinese Massacre had far-reaching effects. It gained national and international attention, and as a result the Chinese consul investigated the affair.47 President Grover Cleveland issued a statement, asserting, “This outrage upon law and treaty engagements was committed by a lawless mob. None of the aggressors, happily for the national good, appear by the reports to have been citizens of the United States.”48 Cleveland went on to quote Article 2 of a treaty between China and the United States that had been signed into law November 17, 1880, stating, “If Chinese laborers or Chinese of any other class, now either permanently or temporarily residing in the territory of the United States, meet with ill-treatment at the hands of any other persons, the Government of the United States will exert all its power to devise measures for their protection.”49 To achieve the terms of this treaty, territorial governor F. E. Warren requested federal troops be deployed from Fort Steele to Rock Springs. The governor’s goal was to ensure that the mines operated normally and that the Chinese miners were protected. Warren’s interests and those of the Union Pacific Railroad and Coal Company were closely related. Trying to relieve any fears the Union Pacific might have, Warren wrote to S. R. Callaway: “Allow me to express my earnest wish that your road will under no circumstances, recede in the slightest degree from the stand taken, that Chinese shall work and that criminals shall not. I believe that future law and good order in this territory as well as discipline of Railroad are at stake. . . . You surely have the United States behind you and China to aid you now.”50 Warren was apparently referring to the presence of the Chinese consul and federal troops sta-

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 279

tioned at Rock Springs. A day earlier Warren sent a telegram from Rock Springs, stating, “Tsang Hay, interpreter, and Wong Sic Chen, Consul at New York, are here under escort of General McCook, conducting an investigation.”51 Warren was pleased with the development. A general and a consul in Rock Springs made him rest easier. He thought the labor problem would finally be resolved and that the incident of September 2 would not be repeated. To show the resolve of the United States to protect the Chinese miners, Camp Pilot Butte was constructed adjacent to where Chinatown had stood. New living quarters for the Chinese were built next to the camp. The army remained in Rock Springs more than a decade (until the Spanish-American War), assigned to protect not only the Chinese but also the railroad and its mines.

Notes 1. Joel Palmer, Journal of Travels over the Rocky Mountains to the Mouth of the Columbia River; Made during the Years 1845 and 1846 (Cincinnati: J. A. and U. P. James, 1847), 35. 2. Interestingly, after “James Bridger” filed his spoliation claim with the Department of War in the 1880s, the then secretary of war William C. Endicott wrote on January 24, 1889, that he would agree to paying Jim Bridger $6,000 for the structures burned in 1857, but he did not want to pay Bridger rent. Endicott claimed that Bridger “did not own or have any title to the land.” William C. Endicott, Secretary of War, War Department, January 24, 1889, manuscript on file, Camp Stambaugh folder, South Pass City State Historical Site, South Pass City, Wyo. (hereafter “SPCSHS”). See also U.S. Congress, Senate, “Report of Committee on Claims regarding the Claim of Jim Bridger against the U.S. Government Involving Ft. Bridger,” May 15, 1892, Missouri Valley Special Collections, Kansas City Public Library, Kansas City, Mo., MV, Q92, B851. 3. From 1990 to 2004 the author excavated the site. 4. A. Dudley Gardner, Data Sets (1990–1991) and Historic Overview (1992) for Fort Bridger (Rock Springs: Western Wyoming Community College and Archaeological Services, 1993); A. Dudley Gardner, David E. Johnson, and Debbie Allen, Data Sets (1992) and Paleoenvironmental Assessment (1993) for Fort Bridger (Rock Springs: Western Wyoming Community College and Archaeological Services, 1994). A. Dudley Gardner, “Traders, Emigrants, and Native Americans, the Mormons and Fort Bridger” (paper presented at the Society for Historical Archaeology 37th Conference on Historical and Underwater Archaeology, St. Louis, 2004). A. Dudley Gardner and Martin Lammers, “Biscuit Root or Biscuits? Trappers and Native Americans at Fort Bonneville, Fort Bridger, and Natural Corrals” (paper presented at the Society for Historical Archaeology 37th Conference on Historical and Underwater Archaeology, St. Louis, 2004). In excavation Level 3 at Fort Bridger, or what is classified as the Mormon occupation, an Elko series point (1990 BC–AD 1080) was recovered. The point appears to have originally been an Elko series point, which was reworked and cut down in size. The point has one of its surfaces partially covered with mortar. More than likely this mortar comes from the construction of the Mormon Wall. Possibly this point was brought in with the gravels used in wall fill or may have been used by the inhabitants as an adaptive reuse of a projectile point they found nearby. 5. Fred R. Gowans and Eugene E. Campbell, Fort Bridger: Island in the Wilderness (Provo, Utah: Brigham Young University Press, 1975).

Fort Bridger and Camps Stambaugh and Pilot Butte

279

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 280

6. Linda Scott, “Pollen and Floral Analysis at Fort Bridger (48UT290), Wyoming” (prepared by PaleoResearch Laboratories, Golden, Colo., April 1991; manuscript on file, Western Wyoming College, Rock Springs), 1– 7; Linda Scott, “Pollen Analysis at Fort Bridger (48UT290), Wyoming” (prepared by PaleoResearch Laboratories, Golden, Colo., April 1992; manuscript on file, Western Wyoming College, Rock Springs), 1– 6; Linda Scott, “Pollen Analysis of Two Features at Fort Bridger (48UT290), Wyoming” (prepared by PaleoResearch Laboratories, Golden, Colo., September 1992; manuscript on file, Western Wyoming College, Rock Springs), 1–5. 7. Gowans and Campbell, Fort Bridger, 43. 8. Scott, “Pollen Analysis of Two Features” (1992), 10. 9. Scott, “Pollen and Floral Analysis,” 10 –11. 10. Ibid. 11. All groups living at Fort Bridger had a variety of guns. For example, a letter from Lewis Robison to D. H. Wells from Fort Bridger, dated April 17, 1856, reads: “I wish to be sure to send those rifles to me the first opportunity. I would like to have you send me one or two camp kettles [also].” The Church of Jesus Christ of Latter-day Saints Archives, Salt Lake City, Fort Bridger file, 2. 12. Scott, “Pollen and Floral Analysis,” 1–11; Scott, “Pollen Analysis” (April 1992); Scott, “Pollen Analysis” (September 1992). 13. Norman F. Furniss, The Mormon Conflict, 1850 –1859 (Westport, Conn.: Greenwood Press, 1977); LeRoy R. Hafen and Ann W. Hafen, eds., The Utah Expedition 1857–1858: A Documentary Account of the United States Military Movement under Colonel Albert Sidney Johnston, and the Resistance by Brigham Young and the Mormon Nauvoo Legion (Glendale, Calif.: Arthur H. Clark, 1982). 14. William Preston Johnston, The Life of General Albert Sidney Johnston Embracing His Services in the Armies of the United States, The Republic of Texas, and the Confederate States (New York: Da Capo Press, 1997), 213. 15. Ibid., 217. 16. Ibid., 214. 17. Ibid., 214, 218. General Fitz-John Porter noted that General Johnston “made constant representations and strenuous efforts to have this route opened.” 18. Fred Gowans and Eugene E. Campbell, The Forts of the Green River Valley (Provo, Utah: Mountain Grizzly Publications, 2003), 99. 19. Hafen and Hafen, The Utah Expedition, 20 –23. Cf. Norman F. Furniss, The Mormon Conflict, 1850 –1859 (Westport, Conn.: Greenwood Press, 1977). 20. Robert S. Ellison, Fort Bridger: A Brief History (Cheyenne: Frontier 280

Printing, 1992), 32 –33. 21. Ellison, Fort Bridger. See Gowans and Campbell, Fort Bridger; Fred R. Gowans and Eugene E. Campbell, Fort Supply: Brigham Young’s Green River Experiment (Provo, Utah: Brigham Young University Press, 1976); William Preston Johnston, The Life of General Albert Sidney Johnston: His Service in the Armies of the United States, the Republic of Texas and the Confederate States (New York: Da Capo Press, 1997, unabridged replication of the 1879 publication). 22. Ellison, Fort Bridger, 48. 23. For horse raids, see Robert A. Murray, “Camp Stambaugh, Wyoming: Guarding the South Pass Gold Fields, 1870 –1878,” 99, 102. See also SPCSHS, Camp Stambaugh file, 4. 24. Murray, “Camp Stambaugh,” 5. 25. John D. McDermott, Dangerous Duty: A History of Frontier Forts in Fremont County Wyoming (Lander, Wyo.: Fremont County Historic Preservation Commission, 1993), 100. 26. “Description and information regarding sale of buildings left standing in 1880 and 1881,” South Pass City State Historical Site, South Pass, Wyoming, Camp Stambaugh file, 4. 27. McDermott, Dangerous Duty, 100 –101. 28. Ibid. 29. Murray, “Camp Stambaugh,” 7. 30. Letter to Major George H. Palmer from Camp Stambaugh to Department of the Platte Received November 17, 1870, SPCSHS, 1. 31. Fred J. Stanton to the Honorable Charles F. Adams, president of the Union Pacific Railroad Company, Boston, September 1, 1884, Union Pacific Railroad Company, MS3761, SG2 President Files, Nebraska State Museum and Archives, Lincoln (hereafter “UPRR”), 1. 32. S. R. Callaway to Charles F. Adams, January 16, 1885, UPRR. 33. Ibid. 34. This quotation was posted in Rock Springs on the night of September 26, 1885, twenty-four days after the Chinese Massacre. Arlen Ray Wilson, “The Rock Springs Chinese Massacre, 1885” (master’s thesis, University of Wyoming, 1967) 58; U.S. Congress, House Reports, 49th Congress, 1885–1886, no. 2044 (Washington, D.C.: Government Printing Office) (hereafter “House Report 2044”), 28. 35. Isaac Bromley, The Chinese Massacre at Rock Springs Wyoming Territory, 1886, 30, in University of Utah Special Collections, Salt Lake City. Isaac Bromley’s book was compiled to present the Union Pacific’s viewpoint of the massacre and is quoted copiously in the company’s History of the Union Pacific Coal Mines (Omaha: Colonial Press, 1940), 46 – 47. 36. Rock Springs Independent, September 3, 1885, contained in House

Dudley Gardner

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 281

Report 2044, 21. This oft-quoted passage is found only in House Report 2044; copies of this or other local papers from the period seem nonexistent. 37. Examples of white attitudes abound; see Murray L. Carrol, “Governor Francis E. Warren, the United States Army and the Chinese Massacre at Rock Springs,” Annals of Wyoming 59 (Fall 1987): Laramie Daily Boomerang, April 2, 1885; Clayton D. Laurie, “Civil Disorder and the Military in Rock Springs, Wyoming: The Army’s Role in the 1885 Chinese Massacre,” Montana: The Magazine of Western History 40 (Summer 1990): 46 – 48. 38. In House Report 2044, 28 –29, the Chinese and white miners differ on this point. The Chinese claimed that the whites became drunk, whereas the Rock Springs Independent mentions nothing of this possibility. 39. Ibid., 28. 40. Salt Lake City Tribune, August 6, 1885, 1. 41. House Report 2004, 29. 42. Salt Lake City Tribune, September 4, 1885, 1. 43. Ibid. 44. Special Orders 103, Headquarter Department of the Platte, Omaha, October 15, 1885, microfilm, U.S. National Archives, Letters Received by the Office of Adjutant General, 1881–1889. See also Salt Lake City Tribune, September 3, 4, 5, 6, and 8, 1885, 1. In the Tribune, stories about the Rock Springs Massacre were carried on page one. 45. U.S. Census 1880, Wyoming. Part of the contradiction found in the census data stems from the official roll sheets being taken from the census records to help in the 1885 investigation of the Chinese Massacre. Investigators used these census rolls to prepare a list of Chinese men living in Rock Springs in 1885. Some of these pages may have been lost. 46. House Report 2044, 28. 47. Toronto Globe, September 4, 1885, 1; Weekly Missoulian, September 11, 1885, 1; Montreal Daily Witness, November 5, 1885, 4. 48. House Report 2044, 16. 49. Ibid., 17. 50. Francis E. Warren to S. R. Callaway, general manager of the Union Pacific Railway, September 18, 1885, microfilm, Territorial Governor’s Communications, National Archives, Denver. 51. Bromley, Chinese Massacre, 87– 88.

Fort Bridger and Camps Stambaugh and Pilot Butte

281

Proulx_06_pp.231-282:Layout 1

8/5/08

12:52 PM

Page 282

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 283

18. Forts Halleck and Fred Steele Annie Proulx

ort Halleck was a small fort with a short life, from 1862 to 1866. It was the only fort between Julesburg and Fort Bridger on the Overland Trail, which in 1862 captured the northern section of the Cherokee Trail. Because of its strategic location, Fort Halleck was vitally important in protecting the new route.1 Major John O’Ferrell of the Eleventh Ohio Volunteer Cavalry commanded the fort briefly until, wrote J. J. Hollingsworth, an enlisted man who was there and who kept a diary, “becoming disgusted at finding ‘graybacks’ on him [, he] resigned and went home to his wife and kids.” His successor, Captain F. W. Shipley, Hollingsworth continued, “robbed the boys out of about $12,000.00 extra duty money.”2 O’Ferrell’s troops built the fort with timber from Elk Mountain’s slopes. By December they had put up barracks for two companies, officers’ quarters, two stables (each of which could hold a hundred horses), a hospital, storehouses, a bakehouse, a sutler’s store, and a jail. The fort’s raisons d’être were to protect the “new” Overland Trail from Indian depredations, to guard Ben Holladay’s stage route over the sixty-five-mile section between the North Platte and Rock Creek stations of the Denver–Salt Lake City run, and to look after the telegraph line and the lonely, dangerous stations along the old South Pass route. They patrolled the trails, sometimes es-

F

corted supply trains and the stages, defended stage stations, even provided horses and drove the stage on occasion. Stages dropped off mail and passengers for the fort at the Elk Mountain stage station, also known as Medicine Bow Crossing, as there was a toll bridge there. Freight wagons unloaded Halleck’s quota of government stores at the same stop. Howard Michael, a stage driver for Holladay, and later a freighter with other companies, wrote that “Rock Creek was a freighting station. All the Government freight was hauled from here to all the surrounding forts.”3 Oscar Collister worked at the Deer Creek telegraph station in 1862. While O’Ferrell’s troops were still building Fort Halleck, the party of Sioux that so alarmed Holladay cut and carried away some sections of telegraph wire at Sweetwater Crossing. Collister and a companion set out on the eighty-mile ride so Collister could cover the station while the regular telegrapher repaired the line. Years later Collister indicated the relatively minor incident of the wire theft was the excuse for the route change: “A few days after I went home it was announced that the stage line was changed to leave Mormon Trail and Julesburg and go to Denver, thence to Bridger’s Pass and Bitter Creek, striking the old line at Green River. The troops remained on the old route to protect the telegraph lines and the emigrants going that way.”4 There was an odd result to the theft of the telegraph wire, a cu283

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 284

rious illness that sounds like an allergic reaction inflamed by imagination and superstition. Collister’s account went on: Some little time after this an important member of the village that the wire-stealing party belonged to, came into Deer Creek with a report that a mysterious disease had appeared in the village and had caused several deaths. The symptoms were a high fever, a rash and soon death followed. The medicine man of the tribe had looked for the cause and found that it originated with a warrior who was wearing wire around his arm, and on further investigation it was found that every member of the party who stole the wire was wearing the same adornment and that several of the party had taken the disease. The medicine man ordered that the wire in camp be buried, and this was done. The disease soon subsided and the belief was universally pronounced among them that the “talking wire” was guarded by the Great Spirit, who avenged the theft and use of it. The telegraph operators were greatly benefited by this calamity, for the next two and a half years the Indians did not molest the wire.5 Troops at Halleck saw considerable action and severe weather. Every winter men froze to death and lost digits and limbs to the surgeon’s frostbite amputations. In late February of 1863 Colonel Collins at Fort Laramie had word that the Utes were running off the horses at the stage stations along the southern route through Bridger Pass. Forty-three men set out to relieve Fort Halleck, Collins following the next day with five more men. The smaller party camped at Rock Creek the intensely cold night of the twentyseventh, and the next morning resumed the march to Halleck just as a storm started. Terrific snow and violent wind made it impossible to start a fire, and the men blundered on, Frank Courtwright falling stiffly from his saddle. The first man to reach Fort Halleck, Private Joseph Hudnell, was frostbitten but survived. Courtwright died a few minutes after he was carried into the fort. Collins wrote in a letter: 284

How poorly prepared they were for that emergency! The little caps, the open-seated saddles which were cool and pleasant for man and horse in warm weather, were of no comfort or use as the wind swept in, chilling the base of the body. One of the caps flew into the air. It had about as much warmth as a bit of dried gourd-shell. A pocket handkerchief was brought out, but benumbed fingers could not manage it, and while they were fumbling with it, it was whisked away and followed the cap. Dismounting, the soldier took off the saddle blanket and attempted to wrap it around his head, but the gale pulled it straight out and it went flying to the skies and was seen no more.6 Dr. John Henry Finfrock (1836 –1893) was the assistant surgeon to the Eleventh Ohio Volunteer Cavalry. He arrived at Fort Halleck on October 10, 1863. After his tour of duty, he settled in Laramie, where he became an important person. He kept a diary while he was at Halleck, recording the weather, injuries, and illnesses of soldiers and emigrants, hunting sorties, and the books he was reading.7 Among his medical cases he lists a man with frozen fingers, an accidental gunshot wound, various fevers, and a boy kicked in the head by a mule. Finfrock picked berries in summer and gathered wild onions in spring, both preventatives of scurvy. His duties were more than medical, for he notes that in May of 1864 he attached a ball and chain to a man. We hear much from historians about frontier justice, and Dr. Finfrock gives us a terse example in his diary. The entry for September 25, 1864, reads: “‘Asa’ killed by boys Co. D—for committing rape upon little girl 12 years old—Dissected him—Exciting time.”8 The familiar tone and a diary entry for August 16 that he “took Asa and went to ‘Rattlesnakes’ for currants” implies that Asa was a man who was with the company, yet not one of them, apparently a servant.9 That he was so summarily dispatched without trial or reflection, dissected by the doctor with whom he had picked berries the month before, a doctor who commented only

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 285

that the grisly chore was “exciting,” hints that he might have been a black man.10 Dr. Finfrock’s brief entry was corroborated sixtythree years later by J. J. Hollingsworth, who stated in 1927 that he remembered little about Fort Halleck “except that nigger who was killed there and whom Ed Lewis, hospital steward, skinned. I met Ed in Laramie last summer and he mentioned that circumstance.”11 Nor was rape and mob execution the only rough action at Fort Halleck. Up in Montana, G. D. Jenks accumulated $4,000 in gold dust. In October of 1864 he stopped at Fort Halleck. A hundred miles beyond the fort he discovered someone had cut his valise and taken his treasure. He went back to Halleck, but although two men, one of them a stage driver, were arrested, both were released and Jenks’s labors in the gold fields were for nothing. Legend has it that the dime-novel writer Ned Buntline came west looking for Bill Comstock, a descendant of James Fenimore Cooper and a rugged frontier interpreter-scout whom Buntline wanted to metamorphose into a character to be named Buffalo Bill. Comstock was killed by Cheyennes before Buntline met him, and the writer’s second choice was William F. Cody. Earlier at Fort Halleck, then under the command of Captain J. L. Humfreville, Comstock had distinguished himself by capturing a defiant murderer, one Bob Jennings, who had ambushed an enemy, then escaped into the wilds of Elk Mountain. Comstock and a dozen Indians tracked Jennings and caught him by saying one of their horses was lame and would Jennings take a look at it and see what he thought. As Jennings bent over the horse’s hoof, the trackers jumped him. He was taken to the post and kept there a while in the guardhouse and then taken out and hung to a large sweep used to swing meat out of the reach of bears and wolves. . . . The large end was so heavy it took two or three men to hold it [the small end] down. When all was ready the captain asked him if he had anything to say. He commenced to say, “Hurrah for Jeff Davis and the Southern confederacy,” but

before he got it all out the Captain said, ‘turn him loose,’ and away he went. The large end of the sweep was so heavy that when it struck the ground it threw him the length of the rope above the pole, and when he dropped down there was just a quiver and all was over.12 The Civil War ended in April 1865, and by June the state volunteers who had enlisted for the duration and who had guarded the telegraph and mail lines were being mustered out and replaced by the last galvanized Yankee regiment—the Sixth U.S. Volunteer Infantry — 950 enlisted men and 26 officers commanded by Colonel Carroll H. Potter. Potter’s territory included the Overland Mail route from Fort Collins to Fort Halleck, a section of the line under almost daily attack. Potter sent out armed escorts with each stagecoach between stations, two of the soldiers a hundred yards ahead of the coach, two more in the rear. And, he added, if the stage line’s stock tenders would keep careful watch during daytime grazing periods and keep the horses in a guarded corral at night, “they will not lose so much of their stock.”13 A few days later sixty horses were stolen from Rock Creek and one soldier killed there. All the stock was taken from Willow Springs and, on the fourth of July, sixteen of the best coach horses were stolen from a heavily guarded corral near Fort Halleck. More raids and attacks hit the eastern part of the line. The various companies of the Sixth were moved around the west like chess pieces. In October 1865, Companies A, D, and F (many of them galvanized Yankees) were transferred from Fort Laramie to cushy Camp Douglas, overlooking Salt Lake City. On the march a rumor swept the troops that they were being sent to the dreadful Camp Douglas prison in Chicago where they had taken the oath of allegiance to the United States only to escape the camp’s horrors, and thirty-five of the men deserted en route, melting into the sagebrush. The remaining men spent the next nine months at the best and pleasantest fort in the west. In June of 1866, Company A fast-marched from Camp Doug-

Forts Halleck and Fred Steele

285

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 286

las to Fort Halleck under command of Major Henry R. Mizner.14 Within weeks Mizner had orders to disassemble some of the buildings, abandon Fort Halleck, and march his troops to the banks of the Big Laramie River seventy miles east, there to build a new fort—Fort John Buford (later Fort Saunders). They were still building in early September when General Sherman came through on an inspection trip and ordered the galvanized Yankees to Fort Kearney for mustering out. Edward Ordway and two companions camped at the old Fort Halleck site after the main buildings were removed. “The old deserted fort was in a very dilapidated condition,” Ordway wrote. “The warehouses and soldiers’ quarters had been built, not in the usual way, but with the logs set endwise in a trench with a very heavy pole and dirt roof, and many of the logs were balsams and had rotted near the ground and by the added weight of snow to their over burden of roof, had fallen down.”15 An 1887 fire destroyed most of the remnant buildings. Fort Fred Steele 16 Massive changes came with the advance of the Union Pacific Railroad—hordes of rough people, large-scale slaughter of game for meat to supply the rail workers, the decimation of forests for ties, rivers filled with floating timbers, and everywhere charcoal kilns, overnight shanty and tent towns. If the railroad made Wyoming, in another sense it destroyed it. In June of 1868, two years after Fort Halleck’s abandonment, Major Richard I. Dodge of the Thirtieth U.S. Infantry, oversaw the building of Fort Fred Steele on the west bank of the North Platte River.17 Fort Steele’s major responsibility was the protection of the railroad and the Overland Trail. The tracks drew ever closer, and then the workers put in a bridge across the North Platte just below the fort. The railroad brought advantages; not only could building materials for the fort be hauled close, but moving troops around became more efficient. There was work for tie hacks, freighters, livery stable operators, stock suppliers, hoteliers, wagonmakers, blacksmiths, and 286

storekeepers. Perhaps rain did not follow the plow, but a kind of coarse-textured American society followed the railroad. The fort was charged with supplying the Ute agency on Milk River and became centrally involved in the longest battle of the Indian Wars, that of Milk Creek, popularly known as the Meeker Massacre. Troops from Fort Steele responded to strikes and riots at the Carbon and Rock Springs coal mines and also sent troops to Rock Springs following the Chinese Massacre there in 1885. And as Fort Steele rose in importance because of its location on the Union Pacific line, famous Fort Laramie, seventy miles north of the railroad, began a long, slow decline.18 Steele had its own official wood and timber reservation. It also served as a timber landing-place for the big pines railroad-cutting crews were taking from the slopes of the surrounding mountains for ties. A log boom stretched across the river near the fort, and a nearby tie hack camp hewed the big trees into ties. Photographs exist that show the uneven, varisized ties so hastily chopped in the race for Promontory Point. The town that grew up near the fort was also known as Fort Steele and served the region’s growing number of timber workers, ranchers and ranch hands, professional hunters and guides, freight haulers, goods drummers, and railroad workers. Fort Steele troops got involved in the battle of Milk Creek on the White River Reservation in Colorado, with high losses, including their commander. Redheaded John F. Finerty was a correspondent for the Chicago Times who had covered the Sioux War and other major stories of the 1870s from Mexico City to the Canadian border. He had been with General Crook at the Battle of the Rosebud in 1876.19 Newspaper correspondents became important in the late nineteenth century, their stories carrying expansionist ideas and justifications to the public. When news of the Ute attack and Thornburgh’s death broke a few days after the fight, Finerty, like every other reporter, had to cover the 1879 battle of Milk Creek ex post facto. There were no reporters on the scene before or at the time of the fight.20 Ever since Nathan Meeker (1817–1879) replaced the old In-

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 287

dian agent, Reverend E. H. Danforth, on the White River Reservation in western Colorado, a simmering hatred had been building among the Utes against the new man. Everything about Meeker, who held the Utes in contempt, seemed wrong. Ohio-born Meeker had worked his way up in the newspaper business. He was widely traveled, an admirer of the French social theorist Charles Fourier, and lived for some years in Trumbull Phalanx, a utopian colony in Ohio. When the Civil War broke out, he became a war correspondent and, later, agricultural editor for Horace Greeley’s New York Tribune, the country’s most influential paper. After the war, still thinking Fourierist ideas, he became the key man in forming the 1870 Union Colony (a.k.a. Greeley) in the South Platte Valley at the junction of the Cache la Poudre and the South Platte, well situated for future irrigation, and a good site for an agricultural settlement. The dictatorial and stubborn Meeker got into financial difficulties and accepted loans from editor Greeley, who inconveniently died in November of 1872. Greeley’s lawyers called in all debts. Even by selling all his colony land, Meeker fell short of the sum owed and was forced to look for a job. After considerable finagling, the job he got was Indian agent at the Ute White River Reservation in western Colorado. It must have seemed to this man, infected with Christian egotism and the hubris that puffed up many men of the period who identified with American expansion and technology, that the reservation was a fresh chance to create an ideal Christian agricultural community. Although he treated the Utes, a tribe that had lived in total independence for centuries on centuries, like his personal vassals, it is possible that he saw the reservation as a tabula rasa on which he could draw his agricultural paradise. Descriptions of Meeker have shown him as a self-righteous reformer, a manipulator and prevaricator contemptuous of the Utes in his charge. His tragedy was that no point of view other than his own had value. The Utes were hunters who controlled the isolated high world of the Southern Rockies, rich in big game because of their

centuries-old stewardship. Their territory was roughly 150,000 square miles of forest, mountain water, and high meadows that extended from northern New Mexico to southern Wyoming. Their hunts for migrating elk and bison took them annually into the Red Desert. They were among the earliest tribes to have horses.21 The Utes were skillful horse stealers, feared by Plains tribes because of their outstanding mounted warriors and their ability to melt into difficult high-elevation terrain that only they knew. They were expert veterinarians and horse breeders; horses were extremely important to them. The Utes had limited contact with whites, even as the Plains tribes struggled with invading hordes. But as the nineteenth century wore on, the whites washed up against the Ute mountains like a rising tide. Eventually, after several treaties, the Ute territory was called a reservation and an Indian agent assigned to them.22 In his zeal to make the Utes into farmers, Meeker shifted the agency headquarters twelve miles southwest to the Powell Valley. Unfortunately, this was a traditional Ute horse-grazing area and the exciting place where they raced their horses. Now this overbearing white man insisted they plow it up, dig irrigation canals, and plant crops. Pitted against the agent were Quinkent, a.k.a. Douglass, so named by an agent who thought he resembled Frederick Douglass, the black abolitionist leader, and three Ute subchiefs — Nicaagat (“Green Leaf” or “he with one earring,” called Ute Jack by the whites), Colorow (a Comanche), and Canella/Canavish, whom the white settlers called Johnson.23 Quinkent, at sixty-one, was the oldest of the three men and, on the surface, the most cooperative with whites on matters of agriculture.24 To the whites, Quinkent/Douglass was the major chief, but to the Utes, who did not recognize leadership pyramids, his influence was minimal. None of the whites involved with the White River Utes noticed that they arranged themselves around lodges belonging to individuals whom they respected. More important than Quinkent was Nicaagat, an Apache-Ute who, as a child, had lived with a Mormon family.25 Before he es-

Forts Halleck and Fred Steele

287

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 288

caped from them he learned to speak English well enough, an ability that made him valuable to both army and reservation agents as an interpreter. He had been to Washington, had interpreted for Crook, knew Thornburgh and many other whites, both military and civilian. He refused to have anything to do with Meeker’s agricultural schemes. Nicaagat was war chief for the White River Utes. Colorow and Nicaagat were openly angry at Meeker’s brusque dismissal of traditional land use, and over the year the tension increased. Meeker began to write letters of complaint to the Department of the Interior and the War Department accusing the Utes of various crimes from raiding stock to setting fires. Major Thomas T. Thornburgh, the commander of Fort Fred Steele, inexperienced at frontier warfare, was filling in his time at Fort Steele, which rumor said would soon close down now that the railroad was finished. Thornburgh received two of Meeker’s letters and checked out the agent’s accusations by contacting local men in a position to know what was happening in the country.26 Thornburgh found the agent’s claims of Indian wrongdoing unsubstantiated. He undoubtedly knew, as did Meeker, that the Utes, unlike the Plains tribes, had never fought with the army.27 Thornburgh was entertaining bankers from the east when a message from General Crook arrived on September 16, ordering him to the White River Agency to respond to the Indian unrest reported by Meeker.28 At Rawlins Thornburgh hired Joe Rankin, who ran the livery stable, to guide the troops to the White River Agency.29 On the morning of September 22, Thornburgh and his troops left Rawlins, on the march to the White River Agency by the wagon route through the Red Desert, the only north-south wagon route at that time.30 On the twenty-fifth they camped on Fortification Creek and Thornburgh sent a letter to Meeker, asking the agent to meet him on the road to report on the current situation. At the agency Meeker read the letter and, perhaps with some spite or braggadocio, told Nicaagat— of all people—that troops were on the way to enforce his policies. Meeker did not ride out to meet 288

Thornburgh, but a concerned Nicaagat did. The Indians waited at Peck’s store on the Yampa for the military to arrive, and stocked up on ammunition. Guide Joe Rankin and Lieutenant Cherry came into the store and persuaded Nicaagat to come with them to Thornburgh. The conversation was limited, Nicaagat wary of troops so near the reservation. The men parted in mutual suspicion. Nicaagat, trying to anticipate all possibilities, set up a guard camp overlooking the agency road. He sent Colorow and several others to meet Thornburgh. An agency employee with a letter from Meeker to Thornburgh traveled with them. This meeting was tense, and Meeker’s letter was vague, saying that “the Indians are greatly excited” and suggesting that Thornburgh come to the agency with only five soldiers, leaving the rest camped outside the reservation. He said nothing about meeting Thornburgh on the road. Because of miscommunication and Meeker’s failure to meet Thornburgh, the tragic drama reached the point of no return. The Utes, highly suspicious and fearing treachery, did not want troops on the reservation, and they let it be known that if troops crossed the boundary line, they would fight. Thornburgh sent a letter back to Meeker, agreeing to come to the agency with five men and a guide. Thornburgh’s second in command, Captain Payne, suggested that the Thornburgh party should go ahead through Coal Canyon, but that he, Payne, and the troops would come down the same route under cover of darkness. Thornburgh, apparently oblivious as to how the Utes might see this, agreed. Both Thornburgh and the troops crossed Milk Creek and entered the reservation. Nicaagat, who had intended to escort Thornburgh and his small party to the agency, was watching from above and dispatched a messenger to Quinkent saying Thornburgh had broken his word. Nicaagat shouted to the warriors not to shoot, but it was too late. The fighting started with Thornburgh and his troops caught in the open. Nicaagat saw Thornburgh fall dead, got up, and walked away, leaving Colorow to continue the fight. He knew, with bitterest clarity, the inevitable consequences.

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 289

Thornburgh’s route to the White River Agency

Minutes before he was killed, agent Meeker wrote a note to Thornburgh telling him, “Things are peaceful,” and saying he would meet Thornburgh the next morning. The message never reached its goal, as the messenger was killed on the way by his Ute guides. Quinkent and his men attacked the agency, killing Meeker and the staff and capturing the women and children, whom they took south to a camp where their own families waited. Guide Joe Rankin left the fight and made his famous ride to Rawlins to summon help, 140 miles in a little over twenty-five

hours. Aid came from Captain Francis Dodge, who commanded Company D, Ninth Cavalry—thirty-five black “buffalo soldiers”— and Colonel Wesley Merritt, commander of Fort D. A. Russell near Cheyenne.31 Afterwards reporter Jerome B. Stillson of the New York Herald, based in Rawlins and suffering from fatal neuritis, sent a legman to the now-cold scene. Finerty and W. P. Boardman of the Denver Tribune rode south with Captain Guy V. Henry of the Third Cavalry, a relief force following Merritt. Everything was over, negoti-

Forts Halleck and Fred Steele

289

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 290

ations with the Utes to get the captive women back were under way, and though public interest was high, after several weeks a disgusted Finerty departed, “while some of the infernally idiotic or knavish benevolists who afflict America with their humane balderdash are trying to patch up an ignominious peace with the murderous Utes.”32 The negotiations for Mrs. Meeker, her daughter Josephine, and Mrs. Price and her children were in the hands of Charles Adams, the former agent at Los Piños and White River. At first the women said they had been treated courteously: “They specifically praised Susan, Ouray’s sister and wife of Chief Johnson, for her kind attention during the twenty-four-day ordeal. No, they had not been harmed or ‘violated’ in any fashion, they responded when specifically questioned by Adams.”33 Later Mrs. Price told Adams a different story, and Adams reported that Mrs. Price, who had been staying with the Meekers, “had been outraged several times during her captivity and she also intimated that Mrs. and Miss Meeker had shared the same fate. Proceeding in my investigation under oath[,] I learned this to be the fact, Chief Douglas[s] [Quinkent] himself having committed the crime on the body of Mrs. Meeker[,] an old lady 65 years old. The ladies are very anxious that for their own sake this fact should be concealed from the public.”34 Adams related the details of the “outrages” to Governor Pitkin “under the seal of confidence.” In short order the news was in the papers and all Colorado was screaming for war against the Utes. The public outcry in Colorado and Wyoming was enormous, thousands of ranting demagogues, newspaper editorials and stories shouting, “The Utes must go!” Loudest and most vitriolic for “extermination” of the Utes was Colorado governor Frederick W. Pitkin. Settlers were poised at the edges of the White River Reservation, waiting to rush in.

290

Notes 1. It was named for a highly regarded military strategist, Major General Henry W. Halleck, a.k.a. “Old Brains,” once a San Francisco attorney and, in April 1862, commander of the Union army. General Halleck had nothing directly to do with Fort Halleck. Brigadier General James Craig picked the name from a list of four worthies. During the Civil War, Halleck opposed the surgeon general’s anguished pleas for a military ambulance corps under the control of the Medical Department rather than the quartermaster. Halleck’s stubborn refusal to consider improved ambulance transport for the wounded undoubtedly contributed to the deaths of many men abandoned to the grisly attentions of conscienceless civilian contractors. See George Worthington Adams, Doctors in Blue (1952; reprint, Louisiana State University Press, Baton Rouge, 1980), 59 – 83. 2. Ibid., 32. Hollingsworth spells the major’s name as “Ferrell.” Oscar Collister renders it “O’Farrel.” Major John O’Ferrell was an officer of the Eleventh Ohio Volunteer Cavalry from January 28, 1862, to April 1, 1865. 3. Howard Michael, “Dictated in August 1925,” Annals of Wyoming 5, nos. 2 –3 (October 1927–January 1928): 80. 4. “Life of Oscar Collister, Wyoming Pioneer, As Told By Himself to Mrs. Chas. Ellis of Difficulty, Wyo.,” Annals of Wyoming 7, no. 1 (July 1930): 357. There was nothing about an attack, Holladay likely counting on the postmaster general’s ignorance of local events. 5. Ibid. 6. Agnes Wright Spring, “‘Surplus Volunteers’ vs. Redskins,” Frontier Times, October–November 1971, 44. 7. An avid reader, perhaps because he suffered from “facial neuralgia” and spent many sleepless nights, Finfrock in his diary entries gives us an idea of the books available at a post in the wilds of Wyoming. Among others, he mentions Fanny Kemble’s Journal of a Residence on a Georgian Plantation, the English actress’s pro-abolition account of her time married to a southern slave-owner, and The Adventures of Gilbert Go-ahead, the novelized version of an interminable serial in the popular children’s paper Robert Merry’s Museum. The book features a Yankee businessman practicing aggressive American commercial interests. A library of about eight hundred donated books at Fort Laramie circulated around the smaller posts. 8. John Henry Finfrock, M.D., Diary, entry for September 25, 1864, American Heritage Center, University of Wyoming. One wonders how many galvanized Yankees of southern origin were in the killing party.

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 291

9. Ibid., August 16 entry. 10. Thomas J. DiLorenzo, The Real Lincoln (Three Rivers Press, New York, 2002), 34, in his discussion of Lincoln’s Emancipation Proclamation comments, “Many slaves who ended up in the hands of the Federal army were not set free but were put to work doing the most unpleasant tasks in and around army encampments.” 11. J. J. Hollingsworth, “Reminiscences,” Annals of Wyoming 5, no. 1 (July 1927): 33. 12. Charles W. Adams, Civil War Reminiscences Interestingly Told (Greenfield, Ohio, n.p., [1917?]), cited in Dee Brown, The Galvanized Yankees (1963; reprint, University of Nebraska Press, Lincoln, 1986), 202 – 203. 13. Official Records of the Union and Confederate Armies, ser. 1, vol. 48, part 2, 921, cited in Brown, Galvanized Yankees, 145. 14. Brown, Galvanized Yankees, 159. 15. Edward Ordway, “Reminiscences of 1866 —,” Annals of Wyoming 5, no. 4 (June 1929): 156. 16. A good review of Fort Fred Steele’s history, with details of its layout and construction, is in Robert A. Murray, “Fort Fred Steele: Desert Outpost on the Union Pacific,” Annals of Wyoming 44, no. 2 (Fall 1972): 139 –206. 17. Colonel Frederick Steele, of the Twentieth U.S. Infantry, had died a few months earlier. A West Point graduate, Steele fought in the war with Mexico, the Civil War, and the Indian wars. Major Dodge should not be confused with General Grenville Dodge, chief engineer for the Union Pacific, nor with Captain Francis Dodge, who came to the rescue of Thornburgh’s troops in the Milk River battle. 18. In later years Fort Laramie was rescued from total oblivion and arson by many interested people, from local Wyomingites, including James Johnston, editor of the Torrington Telegram, and George Hauser, editor of the Guernsey Gazette, to Wyoming, Colorado, and Nebraska historians, landscape historians, wealthy donors, politicians, various early tourist guides, and developers. The National Park Service undertook its restoration from 1958 to 1976 and made it the national historic site it is today. The State of Wyoming saved Fort Bridger from collapse, but no such good fortune came to Fort Fred Steele and other historic sites, despite local champions. Steele’s many buildings are preserved only in photographs, the site today a wasteland of broken glass and crumbled, fire-scorched foundations. 19. By a little twist of fate Ute Jack (Nicaagat), war chief for the White River Utes and a major player in the Meeker affair, had been a scout for

Crook in 1876. Nicaagat also had reputedly rescued Major Guy Henry during the Rosebud fight. A competent English speaker, he lived and died between two worlds. 20. Mark E. Miller, Wyoming state archeologist (and descendant of Isaac Carson Miller, territorial sheriff in Carbon County, 1880 –1884), pieced together military reports and the written accounts of eyewitnesses, locals, survivors, and late newspaper stories in his Hollow Victory: The White River Expedition of 1879 and the Battle of Milk Creek (University Press of Colorado, 1997). 21. Charles S. Marsh, People of the Shining Mountains (Pruett, Boulder, Colo., 1982), 15. The first historical reference to the Utes, according to Charles S. Marsh, was a 1637 description in official Santa Fe records of a battle between the Utes and the Spanish governor’s troops, part of a Spanish campaign to stop Ute horse-stealing raids. Marsh gives no sources beyond assigning the report to “a scribe for Governor Luis de Rosas . . . in the carefully written vellum record books of that day.” 22. The first agent left after one year. His successor, German-born Civil War veteran Charles Adams, was an improvement, thoughtful and concerned for the Utes. Adams was removed when someone back east discovered he was a Catholic, not a Unitarian. His replacement was Nathan Meeker. 23. “Douglass” is often misspelled as “Douglas” in the literature. 24. Meeker paid Quinkent $15 a month. 25. Peter R. Decker, in “The Utes Must Go!” American Expansion and the Removal of a People (Fulcrum, Golden, Colo., 2004), 101, writes: “Captain Jack . . . had grown up orphaned as a mixed blood Apache-Ute in Utah and was later enslaved by a Mormon family. From them he learned to speak English and hate whites before escaping to the White River.” For discussions of the slave trade in the southwest, see L. R. Bailey, Indian Slave Trade in the Southwest (Tower Publications, 1966); and Charles S. Marsh, The Utes of Colorado: People of the Shining Mountains (Pruett Publishing, Boulder, Colo., 1982). 26. The men included Taylor Pennock, W. B. Hugus, J. W. Hugus, and B. T. Ryan. 27. Decker, “The Utes Must Go!” 134. 28. Mark Miller, in Hollow Victory, 13, gives the text of General Crook’s telegram: “Under orders from the General of the Army, you will at once move with sufficient number of troops to arrest such Indian chiefs belonging to White River Agency as are insubordinate, and you will enforce obedience to the requirements of Agent Meeker. You will afford him such protection as the exigency of the case requires and hold the

Forts Halleck and Fred Steele

291

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 292

ringleaders as prisoners until an investigation can be had. You are authorized to suspend orders for movement of ‘E’ Company, third Cavalry, and to use this Company if necessary. Report receipt of this by telegraph and if you require additional troops.” 29. There were three Rankin brothers. Joe ran the livery stable and made a famous ride from the White River Agency to Rawlins after the battle of Milk Creek to report the “massacre.” Robert Rankin was a jailer in Buffalo and Rawlins. James Rankin was a territorial sheriff in Carbon County, elected on the Republican ticket in 1884. M. Wilson Rankin, a cousin of the brothers, was a local rancher who, in his eighties, wrote Reminiscences of Frontier Days, an account of “the Thornburg [sic] and Meeker Massacre” and early life in the Little Snake Valley. 30. Mark Miller describes the route in detail in Hollow Victory, 27ff. 31. During the march the troops averaged sixty miles a day over nearly three days; no horses were lost or damaged; both men and beasts were in good condition and ready to fight when they arrived on the scene. The battle ended with the arrival of Merritt’s troops on October 5 as well as a letter carried from Chief Ouray at Los Piños Indian Agency telling the White River Utes to “cease hostilities against the whites.” 32. Oliver Knight, Following the Indian Wars: The Story of the Newspaper Correspondents among the Indian Campaigners (University of Oklahoma Press, Norman, 1960), 300 –301. 33. Decker, “The Utes Must Go!” 151. 34. Ibid., 155.

292

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 293

19. The Union Pacific Railroad Arrives Annie Proulx

he Union Pacific Railroad, which cut through the center of the Red Desert, created Wyoming. Towns grew up around construction camps and coal mines, then around refueling and water stops. In turn some of these towns became freight and cattle shipping points, sprouting lateral roads and trails into the hinterlands. Early tourism encouraged local people to establish eating houses and hotels. In 1870 thirteen percent of the territory’s labor force worked for the Union Pacific.1 Isolated areas around the perimeter of the Red Desert such as the Little Snake and Sweetwater river valleys connected to the outside world by trails to Rawlins and Wamsutter on the Union Pacific line. The railroad also introduced an endless stream of hoboes and tramps.2 The Union Pacific Railroad pushed through Wyoming in 1867–1868. In the hot summer of 1868 ten thousand men labored across the territory, preparing the roadbed, laying ties and rails.3 Most of them were Union and Confederate veterans of the Civil War. Some were black men. Not a few were Irish. Early in 1867 as the track approached Wyoming, the Indians—fighting for their homeland and still hot with anger over the Sand Creek Massacre of 1864, when Colonel John Chivington’s Third Colorado Volunteers butchered sleeping Cheyennes— stole supplies, attacked survey parties, burned stagecoach stations, and ran off livestock.

T

Percy Browne had the job of surveying the next 275 miles from the Laramie River through the Red Desert to the Green River. He never finished the job. In three days they covered sixty miles across the arid flats of the Red Desert. . . . On the night of July 22 they camped in a dry creek bed that ran into Bitter Creek near the stage station at Laclede. Next morning Browne started down the valley of the creek bed. At about ten o’clock he noticed a large party of Sioux in the distance. . . . Browne and his men galloped toward one of the ridges . . . but the Indians gained too much ground on them. . . . The fight raged for half an hour. Suddenly Browne shrieked in pain, wounded by a ball that passed through his abdomen. . . . Browne died half an hour after reaching the station.4 Two short, tough, well-connected men from Ohio, Union exgeneral Jack Casement and his brother Dan, got the track-laying contract. Men rushed to the work site, hoping for jobs. To solve the logistical problem of laying track across hundreds of miles of empty, townless country, the Casement brothers developed a mov293

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 294

Grading dudley gardner Good grading called for skill. In open country the operator or grader tilted the Fresno handle forward to cut fresh soil.1 The soil peeled off and filled the bucket. The team hauled the dirt to the top of the grade, dumped the load, and evened it out. The choreographed line of horses and graders repeated these steps. Wagons filled with rock and soil also dumped at the end of the grade, as traveling over it tamped down the soils. Bucket by bucket, wagonload by wagonload, a level grade inched over the desert. Clouds of wind-driven dust and grit,

or mud, covered men and animals. But the pressures of the race to reach Promontory Point, the spot where Union Pacific tracks would meet those of the Central Pacific, led to hasty grading, which made a poor roadbed. Along Bitter Creek and other places, grading work required cutting through rock or shale ridges. One such cut, three miles east of Green River City, came to be called Carmichael’s cut. Lewis Carmichael was contracted by the Union Pacific Railroad Company to make some of the most difficult cuts along the grade. Three miles east of Green River an entire camp of Carmichael’s men set up camp as they worked to cut through a solid oil shale bench. Traveler Ed-

able supply depot, the work train, which was a string of specialized cars—materiel, tool storage, smithy, commissary, dormitory, followed by flatcars of ties, rails, bolts, and spikes. Surveyors and locators went ahead of the track along the 42nd parallel, followed by graders and bridge builders. The track layers advanced, throwing down the ties, lifting the rails into place, the gaugers, spikers, and fishplate bolters right behind them with the work train fairly panting at their heels. Every thirty seconds the boss yelled “DOWN!” and a rail dropped into place. The assembly line, credited to Henry Ford, was in full operation decades earlier in the building of the transcontinental railroad. The work rushed forward as quickly as muscle, “bostorial” foremen, and bonus lures could manage. By July the tracklayers were in the Red Desert. That same burning summer, the army was building Fort Steele on the North Platte near the railroad crossing. Fort Steele existed to protect the railroad and its workers, trains and, later, passengers. Terminus town Benton, near Fort Steele, became synonymous with every popular vice of the period. 294

ward Ordway remarked of Carmichael’s graders that they were “a force numbering over 50 men who were natives of all the civilized nations of the earth. Rock men, pick and shovelers and all other necessary helpers in railroad grade making, at that time as a class known as Navvies.”2 Cutting rock required blasting and then the backbreaking work of shoveling the material into a canyon. The transcontinental railroad was literally built by hand. Not only did crews at Carmichael’s camp have to cut through shale, but they also had to build a bridge across Bitter Creek adjacent to the cuts. Carmichael’s camp took on a permanent appearance. A parallel row of homes and tent

J. H. Beadle, the “western correspondent” for the Cincinnati Commercial, wrote a lively book on his travel adventures, The Undeveloped West.5 In the summer heat he left Laramie for justinvented Benton on a construction train, traveling through a portion of the Red Desert: Westward the grassy plain yields rapidly to a desert; at Medicine Bow we took final leave of the last trace of fertility, and traversed a region of alkali flats and red ridges for fifty miles. In the worst part of this desert, just west of the last crossing of the Platte, we found Benton. . . . [N]ot a green tree, shrub or spear of grass was to be seen; the red hills, scorched and bare as if blasted by the lightnings of an angry God, bounded the white basin on the north and east, while to the south and west spread the gray desert till it was interrupted by another range of red and yellow hills. All seemed sacred to the genius of drought and desolation. The whole basin looked as if it might originally have been filled with

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:52 PM

Page 295

pads lined the bank of Bitter Creek north of the cut. Just beyond ran the Overland Trail. And on that trail traveled thousands of freight wagons and hundreds of emigrants and daily stagecoaches, which covered the camp in perpetual dust, each wagon going through the cut and down to a ford across Bitter Creek. Carmichael’s graders blasted rock by day and smoked and drank at night.3 Archaeological excavations conducted in 1986 at Camp Carmichael revealed that, along with champagne and whiskey, the residents consumed bitters by the bottle. Bitters supposedly cured everything from diarrhea to head colds. Based on the number of alcohol bottles

found, it may have been that at Carmichael the bitters eased hangovers. The graders ate well—trout, greater sage-grouse, deer, and beef. Canned goods like peaches and pears and possibly even vegetables added variety to the diet. Fragments of tobacco pipes were found in the excavation. There was something unusual at Carmichael. The crew was cutting through oil shale and did not know it. The depth and length of the cut were useful. The graders put their horses in the cut, built an oily fire at each end, and had a good corral. Horses had to be protected. The usual reason given was that the Indians might steal them. In reality, disgruntled crew members

lye and sand, then dried to the consistency of hard soap, with glistening surface tormenting alike to eye and sense.6 In Benton young Beadle found three thousand residents, streets, blocks of instant buildings purchased from Chicago mailorder houses, twenty-three saloons, half a dozen dance halls. He witnessed fights, woman-beating, drunks, a near-hanging put down by military force, crooked games of chance, dances, and knife play. After months of indulging himself in the delights of Benton, Beadle was nearly broke and had heard nothing from his newspaper about the stories he had sent them. He hired on as a mule whacker for a Mormon freighting company. The route took him through Bridger Pass and across the Red Desert. He described the Bitter Creek region as the “horror of all overland teamsters—where all possible ills of western travel are united.” 7 [B]y nine A.M. the heat was most exhausting. The road was worked up into a bed of blinding white dust by the laborers

who quit the railroad needed a ride home or elsewhere. Relatively few railroad workers had their own stock, so stealing a horse to leave the employment of the railroad was the only choice. One night the fires ignited the shale, and the area became known as the “burning cut.” After the fire was out, the work continued in the fireenlarged cut. 1. Fresnos worked like modern bulldozers and belly dumps but carried only small loads. 2. Edward Ordway, “Reminiscences of 1866,” Annals of Wyoming, no. 4 (June 1929): 157–158. 3. Edward Ordway remarked that he saw the workmen guzzling at the bar of the sutler’s store at Fort Halleck and that they also enjoyed “clog dancing, jumping and pitching horseshoes” (ibid.).

on the railroad grade, and a gray mist of ash and earthy powder hung over the valley, which obscured the sun, but did not lessen its heat. At intervals the “Twenty-Mile Desert,” the “Red Sand Desert,” and the “White Desert” crossed our way, presenting beds of sand and soda, through which the half choked men and animals toiled and struggled, in a dry air and under a scorching sky. In vain the yells and curses of the teamsters doubled and redoubled, blasphemies that one might expect to inspire a mule with diabolical strength; in vain the fearful “black-snake” curled and popped over the animals’ backs, sometimes gashing the skin, and sometimes raising welts the size of one’s finger. For a few rods they would struggle on, dragging the heavy load through the clogging banks, and then stop exhausted, sinking to their knees in the hot and ashy heaps. Then two of us would unite our teams and, with the help of all the rest, drag through to the next piece of solid ground, where, for a few hundred yards the wind had removed the loose

The Union Pacific Railroad Arrives

295

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 296

sand and soda and left bare the flinty and gravelly subsoil. Thus, by the most exhausting labor, we accomplished ten or twelve miles a day. Half an hour or more of temperate coolness then gave us respite till soon after sundown, when the cold wind came down, as if in heavy volumes, from the snowy range, and tropic heat was succeeded by arctic cold with amazing suddenness. On the 27th of August, one of my mules twice fell exhausted with the heat; that night ice formed in our buckets as thick as a pane of glass.8 Ten months later Beadle returned to Benton: “There was not a house or tent to be seen . . . the white dust had covered everything else, and desolation reigned supreme.”9 In May of 1869 the Union Pacific and Central Pacific lines joined at Promontory Point, both Union Pacific’s Dr. Durant, with a screaming hangover, and Central Pacific’s president, Leland Stanford, swung a sledge to drive in the golden spike. Both missed the spike, the ceremony’s anticlimax ignored by the hysterical press. There it was, a railroad line across the entire continent. Hurrah! The haste and skimping with which the railroad had been built now began to show. The bridge over Bitter Creek crumbled away, dumping the engine and express car into the creek and killing one passenger. Unballasted softwood ties in direct contact with soil rotted out, rail joints often did not lie over supporting ties, a number of trestles and bridges had no concrete nor stone support, grades were rough and abrupt. Hundreds of miles of track had to be immediately repaired. Residents in Wyoming later remembered using discarded ties for firewood. A government examiner, Isaac Morris, declared the Union Pacific track “the worst railroad he had ever traveled over, and went on to list its deficiencies: unsafe bridges, tunnels too narrow, roadbed not properly leveled, nor were the rails aligned, roadbed not of the required uniform width of fourteen feet, cross-ties sunken in an unballasted bed. The ties . . . were laid with great irregularity.”10 But the railroad was there, and passengers flocked to make the trip through the Wild West. 296

Notes 1. Lewis L. Gould, Wyoming: A Political History, 1868 –1896 (Yale University Press, New Haven, Conn., 1968), 11. 2. Clark C. Spence writes of the “nationwide tramp epidemic of the 1870s” and on into the 1930s in “Knights of the Tie and Rail—Tramps and Hoboes in the West,” Western Historical Quarterly 11, no. 1 (January 1971). Historians cite the years after the Civil War, alcoholism, migrant crop harvesting, personal reasons, and the depressions of 1873, 1889, and 1929 as causes for men to take to the rails. 3. Ibid., 110. 4. Maury Klein, Union Pacific: The Birth of a Railroad, 1862 –1893 (Doubleday, New York, 1987), 104–105. 5. J. H. Beadle, The Undeveloped West; or, Five Years in the Territories: Being a Complete History of That Vast Region between the Mississippi and the Pacific, Its Resources, Climate, Inhabitants, Natural Curiosities, Etc., Etc. (National Publishing Company, Philadelphia, 1873). The book is still a lively read. 6. Ibid., 87. Beadle’s description is remarkably similar to today’s descriptions of the Red Desert landscape voiced by energy extraction companies. 7. Ibid., 104. 8. Ibid. 9. Ibid., 99. 10. Dee Brown, Hear That Lonesome Whistle Blow: The Epic Story of the Transcontinental Railroads (New York, 1977), 184.

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 297

20. The Union Pacific, the Chinese, and the Japanese Dudley Gardner

Men of the Iron Rails The building of the railroad and the opening of new goldfields altered the ethnic makeup of the mountain west. From dozens of countries, but especially Ireland and China, came railroad laborers. Union Pacific, building in territories with no labor force, had to recruit workers, a need filled by many Civil War veterans and others. The Central Pacific Railroad, building west from Sacramento, used Chinese labor. The Irish workers suited Union Pacific. The building of the transcontinental railroad concentrated Irish people along the string of rails piercing the west. In 1850 a Chinese immigrant passed through Salt Lake City. From 1855 to 1857, Chinese mined gold in Boise Basin. In 1857 a young Chinese boy accompanied trappers to Fort Bridger, Wyoming. New gold discoveries in Colorado in 1858, and in Idaho and Montana in the 1860s, brought in waves of Chinese miners from the California goldfields who contributed their expertise to the development of the Rocky Mountain fields. By the early 1860s, Chinese immigrants could be found from Denver to Montana. Most Chinese figured out that their best chance to make money was to open a business or work as a laborer. Throughout the west they became linked to laundries, restaurants, and railroads. In the mid-1880s a laundry could be opened for less than

$200. In fact, a tub and fire to boil the water put a person in business until a more elaborate operation could be afforded. Chinese cooks knew how to create inexpensive, nutritious meals, and Chinese restaurants became a fixture in the west.1 The Union Pacific in Wyoming employed men from southern China, at first hiring its labor force through tong contacts. By the late nineteenth century the railroad was dealing with Ah Say of Evanston and Rock Springs, a recruiter of Chinese workers. In the Red Desert railroad towns and mining camps, Chinese communities began to emerge By the 1870s, Chinese dominated the section camps, doing most of the hard work of clearing snow in the winter and repairing grades in the summer from the Utah line to Laramie. In the 1870s and 1880s they got their supplies from the Chinatown areas of Rock Springs. Because the railroad went right in front of their houses, they had access to goods shipped from distant places.2 Archaeological excavations indicate that the Chinese in the section camps imported cuttlefish, shark, oysters, and sea bass. They ate out of kaolin ceramic bowls and drank from teacups decorated with four-season designs. They drank whiskey and champagne, smoked tobacco, and consumed opium. More common were herb and bitters bottles, suggesting that the Chinese residents may have used traditional medicines mixed with American medicines 297

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 298

to treat or prevent sickness. In 1869 the Daily Alta California reported that “the Union Pacific people are so satisfied of the superiority of the Chinese for railroad building that they are anxious to secure a large number of them to be employed as repairers and general workmen along their line this summer.”3 Sometime early in 1870 the Union Pacific reached a decision to employ Chinese almost exclusively between Laramie and Ogden, Utah. Union Pacific auditor J. W. Gannett thought the Chinese would labor long and hard enough to enable the company to use fewer repair stations. Fewer section camps meant savings to the company that would make the Union Pacific more efficient. Gannett thought that by using men from Guangdong Province the spacing between camps could be increased from five to six miles. He apparently saw the Irish and the Chinese as labor equals but thought the Chinese workers might work longer hours with fewer complaints.4 The work ethic of the Chinese immigrants made them extremely desirable laborers. Union Pacific found them especially useful in maintaining its main line through southwestern Wyoming. Union Pacific also employed Irish workers, but by 1869 Irishmen began to work in a variety of trades in Carter County, Wyoming. In Carter County, which included parts of today’s Sweetwater, Uinta, and Sublette counties, Irishmen worked as masons, miners, trappers, bartenders, blacksmiths, freighters, mechanics, shoemakers, and glassblowers, the women as cooks and washerwomen.5 Their diverse roles in the local economy show that not long after the railroad had been completed through southwest Wyoming, they made choices to work in service trades and industry. When Irishmen held jobs on the railroad, they were employed as either laborers or foremen. These foremen often oversaw Chinese crews.6 Of the 139 Chinese immigrants living in Wyoming in 1870, 117 worked as railroad laborers. Of the remainder, 2 worked in restaurants, 4 kept house, and 16 washed clothes. The ratio of males to females was 135 to 4. The washerwomen lived and worked with their husbands who were “washer men.” The average age of 298

the workers from Guangdong Province was twenty-three years.7 Although repair crews were needed along the entire length of the railroad in Wyoming, no affordable housing was available to them. Some newly arriving laborers cut homes into creek banks.8 Others, detesting these dugouts, which were vulnerable to flooding, hurriedly made homes from stones or lived in tents. Union Pacific’s solution to the housing dilemma was to construct wooden houses at section camps along its Wyoming main line. These houses were critical shelters in an area that one 1869 observer called the most “forsaken and desolate region I ever saw.”9 One camp in Uinta County, called Aspen, provides insight into these small villages along the railroad. More is known about Aspen because of archaeological excavations conducted there in the summer and fall of 2000. Aspen lies on the west side of the Bear River divide about eighteen railroad miles southeast of Evanston. The beauty of the original railroad line from Carter (north of Fort Bridger) to Evanston was that only one mountain pass had to be crossed to reach the Bear River valley. On the other hand, the harsh Wyoming winters produced snow drifts twenty to thirty feet deep. While the common justification for bringing Chinese into the area was for track repair in good weather, what kept them employed the rest of the year was snow removal. Aspen was segregated. The foreman, railroad agent, and telegraph operator, all permanent employees of the Union Pacific, lived on the west side of the camp. The Chinese and seasonal laborers lived on the east side. In the Chinese section of camp, mixed into the matrix of metal and coal slag, were a wide variety of domestic artifacts, clothing pieces, and food preparation items. Aside from the artifacts uncovered, the diversity of faunal material is one of the most striking features of this site. Li Ping Zhu has argued that when the Chinese immigrated to the United States, one thing that improved was their diet, and, in many cases, so did their general standard of living.10 Not all historians share Zhu’s view and, in at least some cases in North America, Chinese immigrants did experience scurvy and other maladies associated with poor diet.11 However, based on the archaeological discoveries

Dudley Gardner

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 299

at the Aspen site, together with those at Evanston’s Chinatown, it appears that the Chinese immigrants’ diet was not only diverse but also high in nutritional value and calories. Among the floral materials recovered from the late 1800 occupation were raspberries, eggplant, cereals, tomatoes, peppers, carrots, strawberries, currants, gooseberries, plums, cherries, grapes, mint, figs, squash, pumpkin, and olives. Of faunal remains, pigs were the most common species represented, but there were also elements from chickens, cows, goats, sheep, turtles, cuttlefish, oysters, sea bass, and trout. Often labeled as passive immigrants from the Middle Kingdom, the Chinese appeared docile only on the surface. Their roles in opening businesses, forming benevolent associations, and being actively involved in community formation show that Chinese immigrants lived their lives fully. Chinese countered the stereotype of passivity by aggressively developing private businesses and defending themselves from others living on the western frontier.12 Crowded immigrant trains, often with standing room only, carried European and Asian immigrants east and west on the transcontinental railroad. Mrs. Frank Leslie, wife of the owner of the influential Frank Leslie’s Illustrated Newspaper, noted in 1877 on their cross-country train trip that at Green River, Wyoming, she and fellow passengers had “to pass through two or three sleeping cars closely packed, and an emigrant car where, by the dull light, we could see the poor creatures curled and huddled up in heaps for the night, with no possibility of lying down comfortably; but men, women, bundles, baskets, and babies, in one promiscuous heap.”13 Under similar conditions shots rang out in September 1877. In the vicinity of “Granger station, a Chinaman, who was occupying one of the seats in the emigrant car, got up, looked around, took a drink of whiskey, and finally pulled a revolver and commenced shooting at one of the passengers, who was lying behind him.” The first shot entered the right leg “half way between the knee and ankle, shattering the bone.” Making the trip an hour

faster than normal, the train ran full steam to Evanston, where “the wounded man was taken to the Wyoming house and the Chinaman lodged in jail.”14 Other stories report Chinese men involved in frontier brawls. Ah Look was arrested on September 7, 1875, for his “attempt to kill in Evanston.”15 The next May, Look Toy, Ah Toy, and Sing Ten were all arrested for murder in Evanston.16 Four months later Fung Gong received a $250 fine for “assaulting with the intent to murder.”17 The following November, two Chinese men, Ah Hung and Ge Ah Taw, fought one another and were fined a total of $4.18 Revealingly, Fung Gong had made the error of assaulting a white man, so his fine was substantially larger. In Wyoming, as in the rest of the western frontier, gambling consistently caused problems.19 In 1876 in Evanston, Tong Ping, Sing Lin, and Tao Sing were arrested for running a gambling house. Hing Soung, Ah Bic, and Sing Lie also paid fines for selling alcohol and tobacco without a license. In a general sweep to clean up vice in Evanston Chinatown, Ah Moon went to jail and was fined $500 for running a “lewd house.” Those guilty of owning the gaming house paid $300 each to the Uinta County Court, but the unlicensed sellers of liquor forfeited only $2.20 In the Chinatowns of Rock Springs and Evanston, common artifacts found in excavation are gaming pieces and bone dice.21 The railroad connected Chinese workers living in Red Desert settlements with other Chinese communities in Carbon, Rock Springs, and Evanston. The Evanston and Rock Springs communities grew, and Carbon’s waned, with all changes based on the economies of mining and railroading. The Chinese Exclusion Act of 1882 prevented the fresh arrival of new emigrants and workers from China. The 1885 Chinese Massacre in Rock Springs tainted many Chinese people’s view of the opportunities in southwestern Wyoming. Finally, the economic depression of 1893 affected even Wyoming, and some Chinese lost jobs. When the Union Pacific decided to rebuild and realign the railroad in the late nineteenth and early twentieth centuries, they recruited workers from the Land of the Rising Sun.

The Union Pacific, the Chinese, and the Japanese

299

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 300

Japanese railroad workers found their way to Union Pacific section camps early in the twentieth century. Japanese workers, like other immigrants, arrived without families. Many had not yet married. These single males worked in the coal mines and on the railroad, and by 1907 roughly one-third of all coal miners in southwest Wyoming and large numbers of railroad workers were Japanese. Unlike their Chinese counterparts in the nineteenth century, who for a variety of reasons did not bring their wives to America, the Japanese settlers often had relatives arrange marriages, and send their brides to Wyoming. The women undoubtedly suffered a triple shock at arriving in a treeless region, meeting their husbands-to-be for the first time, and discovering they were to live in an isolated section camp in a desert. The Japanese railroad workers isolated themselves from other nationalities. At Point of Rocks the plan map for the camp clearly indicates where the Japanese lived. Yet Union Pacific bosses appreciated their hard work, and over time Japanese became section foremen and lead mechanics. They were highly valued workers and were especially admired for their comprehension of mechanical engineering related to railroad equipment. At Point of Rocks, Thayer Junction, Baxter Station, and Wilkins, excavations of their old house sites turned up Japanese ceramics. Tombstones at places like Thayer, Hanna, Superior, and elsewhere indicate the Japanese retention of Buddhist beliefs in southwest Wyoming. The grave markers also indicate a blending of American and Japanese cultural traditions. Japanese children grew up speaking English and Japanese. They attended local schools and worked in local businesses. In the desert, they played outside their company-built houses and lost toys later found by archaeologists trying to piece together the lives of those who lived in America but whose roots lay in Japan. By 1900 several section camps were made up primarily of Japanese workers. Most of the section camps in Sweetwater County consisted of no less than 40 percent Japanese.22 By 1920 the number of Japanese nationals living in southwestern Wyoming began to decline, but the role they played in maintain300

ing the railroad remained significant.23 Along with Greek section foremen, the Japanese headed railroad crews at several locations in the Red Desert. The Greeks, like the Japanese, began working on the railroad at about the turn of the century.24 George Nishi’s oral interview provides an insight into life in Wyoming for Japanese immigrants: I was born here, but I went to school in Japan. I finished high school in Japan, see. Then I come here, see. Those early days, Japanese figure they must have Japanese educations. So everybody send children to Japan. Like uncle or brother or they take care of them. Only finish so much education they come back to United States. I worked in the mines in 1932 in Blazon. Then let’s see, I worked maybe eight or nine month. That’s about all. but after that this guy used to cut my hair. You know in the mining camp they cut hair; took a bowl and tucked all that in. That saved money, you know. He used to cut my hair all the time and he got killed. After I quit the mines. Name was Okishima. Has got a grave up there. But grave is lost. Early days they make own grave . . . I’m Buddhist. So they give a Buddhist name and that’s put on your tombstone. The—your true name is on there, date died, and date born and your address in Japan. Yoshiye Tonaka relates: My dad first came [here] at the age of around thirteen, fourteen. . . . At the age of thirty-seven he decides he wants to get married, so someone writes him that there is a pretty lady who wants to come to Wyoming because her brother lives in Sheridan. And the brother won’t come back to Japan. This is what my mother says. Of course, my dad was a good looking man. So he goes back to Japan—it’s an arranged marriage—fourteen years difference. So he brings her to Kemmerer and then he says that he’ll take her right to

Dudley Gardner

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 301

Sheridan. Well, he never stopped to tell her that Kemmerer was here and Sheridan was up there. And he was working on the railroad then and so he’d have to come down to Kemmerer to visit my mother. It was 1919 when they were married. She was just going to stay here for a couple of years, and then 1921 the first child was born, 1922 the second [child], then ’26, ’28, and ’30. So she never could go back. Then by that time she was too Americanized. . . . She came to Kemmerer in 1919. Of course it wasn’t too bad then, the buildings were up and everything, but it was just a foreign land to her. She was from a fairly rich family, because she know how to play the samisen and she read the classical book. My dad was orphaned at thirteen, that was it. The older brother was here and sent for him. They were going to make money and go back.25 The period between 1900 and 1940 was marked by fluctuations in the number of Japanese living in the region. This was probably due to the harsh physical environment and to changes in the economic climate of the area. From all accounts, the Japanese of Sweetwater County fared better than the Chinese who preceded them. While racial prejudice was evident, it was of a more subtle variety than the Chinese had experienced in earlier years. Yet despite a relatively stable racial environment, the number of Japanese in the area fell from 436 in 1905 to 187 in 1940.26 The Japanese exit from the railroad camps came swiftly— on February 13, 1942. Immediately after the bombing of Pearl Harbor on December 7, 1941, the Japanese, Italian, and German “aliens” in Rock Springs and the surrounding area were ordered to register with local authorities. The Rock Springs Daily Rocket chronicled this registration, concentrating on the Japanese population in Sweetwater County. Among the first to register were the Japanese miners.27 Only four days after Pearl Harbor, the paper reported, “Officers said the registration had been orderly and that no disturbances had occurred.”28 Two days later the Union Pa-

cific Railroad ordered timekeepers in Rock Springs to “freeze all paychecks of Japanese nationals.”29 Not only the Japanese had to register; Italian and German immigrants were also required to obtain certificates of identification. An article in the Rock Springs Daily Rocket on February 7, 1942, titled “500 Alien Enemies Are Expected to Register in Springs before February 18,” notified people of these nationalities that they would have to carry identification cards with photographs on them.30 A later article, published February 26, claimed that there were not five hundred aliens in Sweetwater County; however, this article was published after the removal of the Japanese.31 On February 13, “all Japanese nationals employed by the Union Pacific Railroad in the area were dismissed.”32 Since the Japanese lived in housing owned by the railroad, termination of their jobs also meant the loss of their homes. The events surrounding the Japanese removal from railroad camps in southwest Wyoming were described by the Rock Springs’ newspapers, such as this account from February 1942: Japanese nationals . . . were given notice to have their belongings and families aboard special cars spotted at sections preparatory to being transported to either Salt Lake City or Cheyenne. They were given three days in which to comply. . . . The sheriff’s office reported that no official orders had been received here for removal or evacuation, but it was understood that the railroad took the step as a precautionary measure. . . . It is not known how many Japanese will be affected by the railroad’s action, but it was stated officially that Japanese nationals not employed by the railroad would not be affected.33 In essence, the Japanese working in the coal mines were allowed to keep their jobs, but those working for the railroad were fired. In the words of one local resident whose family was affected by this action: “My brother was one of them. I was lucky; I was at

The Union Pacific, the Chinese, and the Japanese

301

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 302

school then. But heck, you come home, and they laid off all the citizens, noncitizens, whatever, up and down this line. . . . You were just off the property. You couldn’t even be close to the railroad. Then the railroad comes right through the city [of Rock Springs], and there was [Japanese] people living right alongside the railroad tracks in the city. It didn’t make much sense.”34 During World War II, the Japanese in southwest Wyoming contributed to the war effort by mining and processing coal. At the time, coal was the principal fuel used to fire Union Pacific’s locomotives. These locomotives hauled men and war materials to the Pacific Coast— critical to fight the war in the Pacific. Mixed loyalties were to be expected, but Japanese miners, both male and female, worked in and around the coal mines to provide fuel for the Union Pacific Railroad.35 According to one Japanese coal miner, the only recorded Japanese death in Reliance occurred as a result of an accident in a coal mine. As Yoshio Tabuchi recalls, Tom Kawaguchi “got killed in the tipple. . . . When he fell inside the conveyor, [he went] through . . . the crusher.” Tabuchi’s account represents the only Japanese version of the event that transpired on November 15, 1945. The tipple was a loading and handling facility used to process coal from the Reliance mines and was a relatively modern facility for its day, capable of loading five railroad cars at once. The tipple operator at the time of the death was a twenty-twoyear-old woman who had begun working in the facility after the outbreak of hostilities with Japan. Even though the war with Japan was over when the accident occurred, the wartime operations continued through November of 1945. The woman had attained her position of authority and her skills as a result of the war. Before the war even an Anglo female would not have been employed in a coal mine operation regardless of her skills. In the tipple female workers and high school students gained employment because of the acute labor shortage created by the war. Japanese railroad workers who had been laid off in 1942 had to be replaced. The coal industry was operating at peak production levels; therefore, in 1945, the gender, age, and racial makeup of the workers in the 302

tipple reflected the difficulty in obtaining male laborers. Although this trend of hiring women during World War II has been widely documented, the fact that a Japanese national was working with female laborers in a tipple that provided fuel to the American war transportation network is significant. Life in a company town in Wyoming during the 1930s and early 1940s was not the most desirable. Housing lacked central heating and had been built so hurriedly that occupants complained of snow and dust blown in through the sides of structures. From the dust of the mines to the dust in the streets, coal towns in southwest Wyoming are commonly remembered for the brown and black film that covered everything from bedclothes to cooking utensils. The families settled in the company-built “Japanese Town,” and the bachelor-majority lived in company barracks or at boardinghouses. One Japanese woman took as many as six boarders into her home to help make ends meet.36 She raised seven children in addition to taking care of her husband and boarders. The house was not large, and meals were served at one long table. The Japanese who chose to stay in Rock Springs during World War II either worked in the mines or owned their own businesses. George Okano, a Rock Springs –born resident who eventually fought with the U.S. Army’s famed 442nd in World War II, stated when asked what he did for recreation: “Any days that school was out, by God, guess what? My dad had a store, and you had to help. I can remember, I was just learning how to drive dad’s delivery truck. There used to be a big boney pile down here. In there would be some coal. So dad said, ‘I’ll let you drive that truck, but you go get a load of coal and fill the coal shed up; then you can go hunting or driving or something like that.’ That was recreation, but we had to get the coal first.”37 The Japanese in the coal camps and surrounding communities experienced racial prejudice. George Okano recalls, “As far as [prejudice] goes, it was good. We were real fortunate. But there were some animosities.” Chinese residents in Rock Springs “used to run around saying, ‘I’m Chinese.’ One of them kicked Jeral,

Dudley Gardner

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 303

my brother, out of their restaurant here in town. I never did forgive them. I won’t eat there even today.”38 Edith Sunada claims, “People weren’t very kind. In fact, we had one man come in one night . . . drunk, and he kicked the door down . . . trying to get in. I don’t know what he thought he was trying to do.”39 Union Pacific filled the Japanese worker vacancies they had created on February 13, 1942, by hiring laborers of Hispanic/ Latino descent and, later, Navajo repairmen. “Mexican” and “New Mexican” railroad workers began to appear in the census records early in the twentieth century. In 1910 the U.S. Census enumerators at Wamsutter noted Latino and Greek railroad workers working alongside Japanese men. The Union Pacific would eventually use Greek and Japanese foremen to supervise Hispanic crews. At Creston in 1920, one entire repair crew came from Mexico; their foreman was born in Japan.40 In 1942 when the Union Pacific Railroad Company fired the Japanese railroad workers, they were firing the lead men from the crews. Mexican workers, already present at most section camps, replaced the Japanese foremen. Most of the section camps had to use New Mexican and Mexican workers until Union Pacific closed the camps in the 1980s. The use of Mexican railroad workers continues to the present. From Irish immigrants to Mexican laborers, Union Pacific followed the same path—the use of hardworking minorities to staff and repair the railroad line across southwestern Wyoming. This policy carried an attractive financial benefit—paying the emigrants or minority laborers less than other workers. Living in isolated camps along parallel ribbons of steel, these laborers brought ethnic and cultural diversity to the Red Desert.

Notes 1. Frank Leslie, writing for his Frank Leslie’s Illustrated Newspaper on an 1877 railroad trip from New York to San Francisco, delightedly described the Chinese waiters at the highly regarded Mountain Front Hotel in Evanston, commenting that he was thrilled to see “serene moonfaces and almond eyes . . . in place of the unshorn, unwashed, unsavory Celtic countenance.” Richard Reinhardt, Out West on the Overland Trail: Across the Continent Excursion with Leslie’s Magazine in 1877 and the Overland Trip in 1967 (American West Publishing Co., 1967), 86. 2. In the preconstruction railroad dreams that fired both the public and promoters was the idea that a transcontinental railroad would open Asian markets. Collis Huntington, the aggressive power behind the Central Pacific, envisioned a fleet of ships bringing silks, spices, and teas from the Orient to San Francisco, where the luxury goods could be loaded onto “Silk and Tea Trains” and shipped to eastern American markets at great profit. In the 1870s some silks and tea were shipped east by rail, but the trade faltered as the Suez Canal, finished only a few months after the transcontinental railroad, brought shipping costs lower than any railroad could. Yet the connection allowed Chinese laborers living in the remote Red Desert to enjoy traditional delicacies and cultural amenities. 3. Daily Alta California, May 8, 1869, 1. 4. J. W. Gannett to Oliver Wendel Ames, August 27, 1870, Union Pacific Railroad Company Collection, Lincoln, S62, Box 6, Office of the President, Nebraska State Museums and Archives, 1–2. 5. Special Census of Wyoming Territory, 1869. 6. Ibid.; Ninth Census of the United States 1870 (Government Printing Office, Washington, D.C., 1872), Wyoming. 7. United States Census 1870, Wyoming. 8. A general description of this scene appears in William A. Bell, “The Pacific Railroads,” Fortnightly Review 5 (January–June 1869), 572 –573. 9. Bell, “Pacific Railroads,” 570. 10. Li Ping Zhu, A Chinaman’s Chance: The Chinese on the Rocky Mountain Frontier (University Press of Colorado, Niwot, 1997). 11. A. Dudley Gardner, “Two Paths, One Destiny: A Comparison of Chinese Communities in Alberta, British Columbia, Montana and Wyoming, 1848 –1910” (doctoral diss., University of New Mexico, Albuquerque, 2000). 12. Cheyenne Daily Leader, September 28, 1877, 2; Helena Herald, “Supplement,” June 26, 1867, 1; Weekly Missoulian, June 17, 1881, 1; July 22, 1881, 4. The archeological record shows that the residents of Chinatowns

The Union Pacific, the Chinese, and the Japanese

303

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 304

were well armed. These records demonstrate that most Chinatowns contained bullets and gun parts. See, for example, David L. Felton, Frank Lortie, and Peter D. Schulz, The Chinese Laundry on Second Street: Papers on Archeology at the Woodland Opera House Site, Sacramento Department of Parks and Recreation, California Archeological Reports, no. 24 (1984), 76 –77. The Evanston Chinatown and Tucson Chinatowns also contained firearm parts and ammunition in the Chinese occupation horizons. J. Homer Thiel, Archeological Investigations of a Chinese Gardener’s Household (Center for Desert Archeology, Tucson, Ariz., 1996). 13. Richard Reinhardt, Out West on the Overland Train: Across-theContinent in 1877 and the Overland Trip in 1967 (American West Publishing Co., Palo Alto, Calif., 1968), 80. 14. Cheyenne Daily Leader, September 28, 1877, 2. 15. Book 1, Jail Record, January 4, 1875–May 24, 1904, Uinta County Sheriff’s Office, Evanston, Wyo. (hereafter “Book 1, Jail Record“), entry 85, 3. 16. Ibid., May 12, 1876, entries 136, 138, 139, 4–5. 17. Ibid., September 22, 1876, entry 113, 10. 18. Ibid., November 2, 1876, entries 188, 189, 6 –7. See also the Helena Weekly Herald, March 19, 1868, 7; and Daily Alta California, May 8, 1869, 1. 19. Helena Weekly Herald, September 19, 1868, 8; King v. Yung Sam, Ah Sam testimony, March 7, 1865, Call no. GR 149, box 4, file 1865/15, Provincial Archives, British Columbia, Victoria, 1–2. 20. Book 1, Jail Record, entries 151–157, 4–5. 21. Kevin W. Thompson, A. Dudley Gardner, and Russel Tanner, “Archeological Investigations at the Rock Springs Chinatown,” Small Report and Article File (Archeological Services of Western Wyoming Community College, Rock Springs, 1992), 1–10; and A. Dudley Gardner, “Results of Excavations at the Evanston Chinatown,” Small Reports and Article File (Archeological Services of Western Wyoming Community college, Rock Springs, 1996), 1–5. 22. Twelfth Census of the United States 1900 (U.S. Government Printing Office, Washington, D.C., 1901), Wyoming. 23. Fourteenth Census of the United States 1920 (U.S. Government Printing Office, Washington, D.C., 1922), Wyoming. 24. Eleventh Census of the United States 1910 (U.S. Government Printing Office, Washington, D.C., 1913), Wyoming. 25. Yoshiye Tanaka, interview with author, Kemmerer, Wyo., 1991 (manuscript on file, Archeological Services of Western Wyoming College, Rock Springs). See also Yoshiye Tanaka, “No Brand Name: Home-

304

made” (manuscript on file, Archeological Services of Western Wyoming College, Rock Springs), 1–11; A. Dudley Gardner, “World War II and the Japanese Americans of Southwest Wyoming,” Wyoming History Journal 68, no. 2 (Spring 1996): 22 –32. 26. Sixteenth Census of the United States 1940 (U.S. Government Printing Office, Washington, D.C., 1942), Wyoming. 27. Rock Springs Daily Rocket, December 10, 1941, 1. 28. Ibid., December 11, 1941, 1. 29. Ibid., December 13, 1941, 1. 30. Ibid., February 7, 1942, 1. 31. Ibid., February 26, 1942, 1. 32. Ibid., February 13, 1942, 1. 33. Ibid. 34. George Okano, interview with Dudley Gardner, Rock Springs, Wyo., 1984. (Manuscript on file, Western Wyoming Community College, Rock Springs), 6. 35. Agnes Sunada Tabuchi and Yoshio Tabuchi, interview with author, Reliance, Wyo., 1986 (manuscript on file, Archeological Services of Western Wyoming College, Rock Springs). Yoshio Tabuchi passed away at the age of seventy on February 7, 1988. From the Rock Springs Daily RocketMiner, February 9, 1988, 11: “He was born February 28, married Agnes Sunada on October 14, 1940, in Manila, Utah. . . . He worked as a coal miner for 22 years.” 36. Agnes and Yoshio Tabuchi interview, 16 –17. 37. George Okano interview, 3–5. 38. Ibid., 6. 39. Edith Sunada interview, 26. 40. Eleventh Census 1910, Wyoming; Twelfth Census 1920, Wyoming.

Dudley Gardner

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 305

21. Inhabitants of the Margins Annie Proulx

he larger Red Desert is a series of basins and badlands called by an interlocking confusion of names—Great Divide basin, Red Desert basin, Wyoming basin, Green River basin, Washakie basin, Sand Wash basin, Cherokee basin, Poison basin, and Baxter basin, demarcated by scores of rims, ridges, and buttes. Most of the area is sagebrush steppe rather than desert.1 How we call it is a matter of perspective. The 1850 emigrants from the wooded east trudging through the alkali dust, longing for shade and water, knew they were in a cruel desert. But an Egyptian scientist, familiar with the Sahara, remarked when visiting Nevada’s Great Basin, whose natural history features are similar to those of the Red Desert, that what he saw was “not a desert but a veritable botanical garden.”2 It is convenient to divide the desert into north and south regions, I-80 serving as the demarcation. The outstanding features of the north half are the Boar’s Tusk, Jack Morrow Hills, the Honeycombs, Oregon Buttes, White Mountain, the CaliforniaOregon-Mormon Trail, South Pass. The southern section in the Washakie basin has Adobe Town, Kinney Rim, Prehistoric and Skull rims, Man and Boy buttes, the Haystack Mountains, the old Cherokee Trail. The desert is not pristine. It is dotted and crisscrossed with pipelines, power lines, stone cairns, thousands of miles of rough

T

roads, new roads constantly added, transmission towers, stock tanks, airstrips, the remains of horse traps and juniper corrals, and the ruins of old stage stations and ranches. The eye ignores most of these, and it is the absence of fences and graceless railroad town architecture, as well as the long sight lines that let us believe we are in a wild landscape that has changed little over the centuries. Today the greater Red Desert is chopped up by hundreds of new energy extraction roads. Heavy truck traffic and compression machinery invade the desert’s former silence. Ancient cultural sites, many still unexamined, are being destroyed for lack of cultural resource protection.3 Some energy companies work hard to reduce their impact on the ecosystem. Some do not. Once various Indian tribes counted this as their territory. These people maintained stable cultures for thousands of years, traversing and living in the basins, hunting bison, elk, pronghorn, deer, and wild horses. Today we still find their stone hunting blinds, arrowheads, petroglyphs, and rock paintings throughout the area. Fur traders and explorers followed the advice of Indians concerning routes through the mountains. Their other guides were trial and error, game trails— especially buffalo roads “like so many well-traveled highways”—streams and rivers, ridges, and distant landmarks.4 By the 1840s, when the beaver market collapsed and 305

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 306

the company fur trappers’ days in the wilderness ended (a few solo trappers continued to take furs for decades), emigrants began to use what became the great passageway to the golden west, the Wyoming portion of the old Indian Road which became incorporated into the California-Oregon-Mormon Trail. This 2,100-milelong road sawed across the country and through the western tribes’ ancient world. The whites’ rhetoric that a distant Great White Father protected the road and its white travelers led the Lakota Sioux to work out the idea that the road itself, as place, must hold some sacred power for whites. They called it “the Holy Road,” but not with reverence.5 This trail ran along the northern edge of the Red Desert.6 Historians have estimated that between 250,000 and 500,000 people made the arduous eight-month journey from the Missouri River to the west coast during the decades from the 1840s to the 1870s, one of the world’s great human migrations. Seven of ten emigrants were farmers, almost none from New England or the coastal south, but from the midwestern states of Iowa, Illinois, Indiana, and Missouri. Most of them were poor. Farming then was not a business but the ordinary family’s life calling. Many of these farmers commented in their diaries on the fantastic vistas and scenery they saw while crossing the Red Desert, but they did not pull out of the wagon trains to settle Wyoming.7 They aimed to get their hands on rich soil; the arid sagebrush steppe and alkali flats were hardly agricultural temptations. The farmers could not get away from the parched desert basins soon enough and, in their journals, expressed misery in, and loathing for, the Red Desert, as well as a fear of Indians, even as they goggled at the awesome landscape furniture. Their discomforts drove them to exasperation; merciless sun, no shade, jolting wagons, thirsty and footsore oxen, sulphur water, constant wind. That infamous wind carried fine alkali dust, highly soluble in water, so that moist eyes, licked lips, or a wet cloth swabbed over a hot face produced a caustic burning sensation. Women complained of sore and irritated lips. Nerves wore raw and tempers short. 306

Soldiers at the posts and forts strung along the trail found the emigrants a nuisance. The obsessive search for grass and water for stock was foremost in the minds of travelers as well as troops on the move. This need for forage and water was a constant worry in the nineteenth century, keener than we can now imagine. Battles and lives were lost through wrong decisions on where to camp. Corporal Hervey Johnson of Company G, First Battalion of Ohio Sixth Regiment Volunteers, whose job it was to protect emigrants and the new telegraph wire, served in the 1860s at Fort Laramie, later at the small Pacific Springs telegraph station near South Pass, and finally at Deer Creek on the North Platte. He wrote: We have been bothered to vexation almost by importunate questions of the emigrants. A hundred men will pass in a day all will ask the same questions such as how far is it to grass? Any wood there? Is the road sandy? How far is it to the crossing how far is it to the bridge? What is the toll. How far is it to Bannac how far to fort Bridger, how far to Salt Lake, is there a post office here? . . . [S]ome one will ask how far it is to grass, then ask half a dozen questions about the road and other things, then ask how far it is to feed, then a dozen more questions, then ask if there is any grass up there. More questions, then is the feed good up there, and so on.8 Many of the early settlers around the edges of the Red Desert were ex-army men who had come into the country on the military’s ticket, or veterans of the Civil War bumming around the west, as well as railroad workers and stagecoach personnel who had learned to understand and appreciate the terrain in the course of their duties. Joel J. Hurt, in 1865 a guard on Holladay’s stage line, became one of Wyoming’s premier early sheepmen. Tim Kinney, a railroad worker, settled in Rock Springs, got into the cattle business, and later switched to sheep. Of these men, not a few came from green and misty Ireland and perhaps had experi-

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 307

enced the crowded “paddy shack” slums of New York and Boston before joining up and coming west. The Irish in particular, with little chance to own land back home, saw that here were immense tracts, free to the hardworking homesteader. The fringes of the desert and even the interior looked like worthless land to farmers, but these men were not thinking of farms. Herds of cattle and sheep would fill up the empty space recently occupied by bison and Indians. There was a real chance—if you could seize it—to build a personal kingdom. And there was that endless, brilliant sunlight. The Irish and, to some extent, the Scots, went in for sheep, while the English favored cattle. Not only a difference in investment sums needed to start up but a kind of social snobbishness seemed at work here. The cattle industry attracted the British upper classes and nobility who met at the elegant Cheyenne Club, where rank, money, and connections indicated an individual’s worth. Gradually people settled around the basin rims —Rawlins, Wamsutter, Hanna, Carbon, Superior, and Rock Springs along the Wamsutter Arch; ranches in the Sweetwater and North Platte valleys; Savery, Dixon, and Baggs in the Little Snake River valley; Brown’s Hill and Tullis, where the desert laps against the west flank of the Sierra Madres; Slater, Steamboat Springs, Lay, Lily Park, Hayden, and Craig in northwest Colorado; and Vernal, Utah, as well as Brown’s Hole near the junction of Utah, Wyoming, and Colorado. These early towns developed very different characters, and around them dozens of tiny settlements appeared briefly like candles in the dark, then vanished. Raising livestock— cattle, sheep, horses—was the way most Anglos made a living in early Wyoming. But there were a few other occupations and a few other ethnic groups. As the Union Pacific Railroad developed the coal seams along its route, it hired large numbers of foreigners to work in the mines —Chinese, Finns, Japanese, Welsh, Swedes. Moreover, the railroad advertised heavily in newspapers and brochures, offering for sale the rich plunder of land awarded it by the federal government in a twenty-

mile-wide checkerboard distribution along each side of the route. The railroads even sent representatives to foreign countries, their job to persuade land-starved peasants to buy railroad property in the American west. As enticements, they sometimes offered free seed, often of poor quality and loaded with exotic weeds whose descendants continue to plague the region. Homesteading was difficult, described in R. H. Burns’s Wyoming’s Pioneer Ranches as composed of “heartaches, denial and hard work.”9 People lived, struggled, moved out, tried again, shifted to town and finally died there. Their years in and on the Red Desert are mostly lost now. The keys to success in this harsh place were a supporting share of the raw resources, transportation, cash flow, and cooperation with neighbors. Lives depended on that cooperation. As population increased, the competition for good stock range sharpened and caused sheep-cattle “wars” and the drawing up of boundary lines that were sometimes literally “dead lines.” Roads ranged from bad to terrible, and getting in and out of a remote ranch hung entirely on weather. Bentonite in the Red Desert soils swelled and held water, creating dry-weather-only roads that became greasy and impossible to move over when it rained. The sticky mud built up on wagon wheels, horses’ hooves, and the shoes of the human walker. Winter storms closed roads for weeks, even months. As for cash flow, there was none. Cattlemen and sheepmen borrowed money against their ranches and livestock, made it until autumn when the stock was shipped, paid back the loans and interest, and, more than likely, borrowed again to make it through the next year. Small ranchmen worked for others for wages, trapped, freighted, worked for the army posts, and, in 1867 and 1868, labored on the railroad. In the twentieth century the Great Depression of the 1930s did not hit rural people as hard as those in towns and cities. Not much was different. People slipped into a barter economy. Vegetable gardens, venison, beef or mutton, the home dairy cow, a few pigs and chickens, plus still-abundant game supplied the ba-

Inhabitants of the Margins

307

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 308

sics of life. The horse continued to be useful for transportation. Flour, tobacco, sugar, coffee, and baking powder had to be paid for in cash. The sources of cash for most people in the settlements around the Red Desert were ranch work for wages, wild-horse catching, and the making of bootleg whiskey. The few women— often teachers— could only hope to graduate from their positions in one-room schools to marriage and the exalted stations of ranchers’ wives. One of the submyths of the west holds that prostitutes married ranchers and cowboys and moved out of the profession, but the work of historian Anne M. Butler, in her Daughters of Joy, showed that most men chose mates from their original hometowns, not from the whorehouses, and that those prostitutes who did marry ended up with gamblers, pimps, and troubled transients whose problems and miseries equaled their own.10 Major William Shepherd, a sharp-eyed Englishman who drove a band of sheep from California to Wyoming, remarked, “The great want in the prairies is the right kind of girls.”11 Life around the edges of the Red Desert continued relatively unchanged for more than a century, even when oil wells were punched into the scrubland, with people in the tiny settlements doing their major shopping at regional centers such as Rawlins, Rock Springs, and Craig, Colorado.12 The completion of I-80 in the 1970s, a major cross-country four-lane highway that cut the Red Desert in half and prompted travelers looking out at endless sagebrush to wonder why anyone wanted to live in such a desolate place, made it easier for rim dwellers to get to towns for medical services and supplies. Terrific winds and blowing snow still make the section of I-80 that runs through the Red Desert extremely hazardous in winter. It is not unusual for miles of the highway to be closed for days at a time in winter, or to see huge semis on their sides at the margin of the road, blown over by violent wind gusts. In 1977 an entity calling itself the Triad Land Company, owned by one James Butler and a Palm Springs, California, couple, Merle and Patricia Brooks, ran an ad in Field and Stream magazine of308

fering forty-acre “ranches” for $50 down and $45 a month. The land was in the Red Desert, part of the abandoned Barrel Springs ranch, roughly twenty-five miles south of Wamsutter. The only road was little more than a rough trail. There was no electricity, no water, no town, no school bus service. The dust-laden wind blew constantly. A ten-page “fact sheet” from the Palm Springs couple told would-be buyers that “seeding and watering will produce good hay fields, gardens, trees and shrubs, although irrigation is essential to raising a good crop.”13 Federal officials scoffed at the idea of such greenery growing in a desert that receives seven to nine inches of rain annually. Despite the huge obstacles, eighteen families from various parts of the country bought land. Their forty-acre parcels were widely separated, each remote and without the benefit of neighbors. During the summer of 1977 they put up cabins and tarpaper shacks and tried to find jobs in the distant towns. Some local people saw them not as brave pioneers but as fools; others were concerned that they would all freeze to death during the savage ground blizzards that characterize the desert. One “pioneer” job seeker was told, “We don’t hire animals.”14 In the end the Barrel Springs pioneers moved on to unknown parts. The land scams continue today with offerings on eBay. Although the Red Desert demanded great fortitude from those who chose it as a place to make a living and a home, many took up the challenge and established home ranches on the desert.

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 309

Notes 1. See George P. Jones, “Sagebrush,” this volume. 2. Stephen Trimble, The Sagebrush Ocean (University of Nevada Press, Reno, 1989), 19. 3. In May 2006 the National Trust for Historic Preservation released a report criticizing the Bureau of Land Management’s inability to protect thousands of archeological and cultural sites ranging from petroglyphs to cliff dwellings. 4. Washington Irving, Astoria (John B. Alden, New York, 1887), 141. 5. Peter Nabokov, “Orientations from Their Side: Dimensions of Native American Cartographic Discourse,” in G. Malcolm Lewis, Cartographic Encounters: Perspectives on Native American Mapmaking and Map Use (University of Chicago Press, Chicago, 1998), 244. 6. J. V. Frederick, Ben Holladay, the Stagecoach King (reprint, University of Nebraska Press, Lincoln, 1989), 260. The more southerly route, incorporating a section of the 1850 Cherokee Trail, Holladay put in place in 1862 after apparent Indian trouble west of Fort Laramie; it is also called the Overland Trail. To avoid confusion, we call the older route the California-Oregon-Mormon Trail, and the short-lived, safer southern route the Overland Trail, which was used until the Union Pacific began carrying passengers in 1868. In 1866 the astute Holladay, who knew the railroad was coming, sold out to Wells Fargo. 7. John Mack Faragher, Women and Men on the Overland Trail, 2nd ed. (Yale University Press, 2000). Faragher notes that emigrant diaries contained three major themes: “practical matters, health and safety, and natural beauty” (14). 8. William E. Unrau, Tending the Talking Wire: A Buck Soldier’s View of Indian Country 1863–1866 (University of Utah Press, Salt Lake City, 1979), 162 –163. The soldiers made sport of the emigrants, who struck them as gape-mouthed ignoramuses who would believe anything told them. Corporal Johnson caught an eagle, and before it died, the bird was a subject of curiosity to passing emigrants. He wrote: “It [the eagle] staid around a good while and was the wonder of all emigrants, a great many of whom did not know what it was. They would always ask a great many foolish questions about it and other curious things they saw, and always got as foolish answers. One day I was sitting out under the shade of the arbor, not interrupting any body, and an emigrant came staring up and looking at the bird awhile said ‘Mister what kind of a bird do you call that there’ I told him that was the bird of liberty, that roosted on the liberty pole, and crowed, and sung yankee doodle and hail Columbia. He went off satisfied with the information.”

9. R. H. Burns, Wyoming’s Pioneer Ranches (Top of the World Press, Laramie, 1955), 607. One of the best books on hardscrabble ranching in the Red Desert fringe region is John Rodney Barney, Looking Back: My True Memories from 1908 –1996 (Commerce Printing, Sacramento, Calif., 1999). 10. Anne M. Butler, Daughters of Joy, Sisters of Misery: Prostitutes in the American West 1865 –90 (University of Illinois Press, Urbana, 1987), 52, 151. 11. Major R. E. William Shepherd, Prairie Experiences in Handling Cattle and Sheep (O. Judd Co., New York, 1885), 131. 12. For a lively description of Rawlins in 1950, see John Niland, A History of Sheep Raising in the Great Divide Basin of Wyoming (Lagumo Corp., Cheyenne, 1994), 147–163. 13. Howard Pankratz, Denver Post, November 13, 1977. 14. Michael Coakley, Chicago Sunday Tribune, November 13, 1977.

Inhabitants of the Margins

309

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 310

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 311

22. The Little Snake River Valley Annie Proulx

he mountain men and trappers working the Yampa, the Bear, Elk Head, and streams east of the Green were in the Little Snake River valley fifty years before the first settlers.1 The valley was lush in comparison with the surrounding desert, and thick with game. A favorite Indian gathering site for annual trade fairs lay northeast on the Encampment River. Trappertraders found it a convenient place to swap goods for furs. Fraeb’s Post, established in 1837 or 1838 by Henry Fraeb and Jim Bridger, stood briefly on what was then called the St. Vrain’s Fork of the Elkhead River. The streams were later renamed Savery Creek and the Little Snake River, as they are known today. A few prospectors drifted in, found a little gold. In the 1870s a handful of trappers and some of the miners settled in the valley and began to build up small herds of cattle and horses. More settlers came, despite the looming presence of the Red Desert. In the valley some picked out choice river bottomland with deep soil and good water. Others retreated to isolated ranches where they lived hard and elemental lives. A few gravitated to the most inaccessible corners, which became known as outlaw refuges. The upper and lower Powder Springs (sometimes called the Powders) were close to the Wyoming-Colorado line, and the upper Powder Springs and Dripping Springs to the east served different gangs as outlaw resorts. In time the whole valley of the

T

Little Snake developed a reputation for friendliness to outlaws.2 In 1870 two sometime miners, Bill Slater and John Brockmeyer (or Brockmire, as his tombstone has it), the latter better known as Bibleback Brown, built a cabin at the confluence of Savery Creek and the Little Snake, the first permanent structure in the valley. Later Savery, Dixon, and Baggs became the principal settlements on the north bank of the Little Snake. The cattle business expanded. In 1877 there were 5,000 cows in Carbon County. Three years later there were 72,055.3 Although the relations between small-scale ranchers and the big operators never reached the pitch of animosity as in Wyoming’s Johnson County, the old accounts show there was friction within the cattle club, especially as the range became overgrazed. In 1890, Wyoming Territory became a state. Signs of trouble began to show in the Little Snake Valley. The major problems in the 1890s, residents thought, were sheep, lingering Utes, and game laws. For those in northwest Colorado, the fact of Wyoming as a state seemed as intrusive as if a foreign power had seized the anything-goes territory once open to all. Cattle graziers of northwest Colorado disliked the Wyoming sheep flocks that pressed across the Little Snake onto the public land locals considered their personal cow range. The cattlemen 311

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 312

and their cowboys reacted with violence against the encroaching sheep, sheepherders, and sheep owners. The Little Snake River became the cattle-sheep deadline—sheep to the Red Desert north of the river, cows to the south, and never the twain should meet. Colorado animosity targeted two gritty Welsh brothers, Jack and Griff Edwards, tutored in stubbornness and staying power in Australia and Brown’s Hole. They pushed their sheep flocks south across the Little Snake into forbidden territory.4 In 1896, after promising Colorado cattlemen that he would keep his sheep off “their” range, Jack Edwards broke that promise. In revenge, three hundred sheep were clubbed and two of his sheepherders shot dead, the third allowed to escape to decoy Edwards to the scene. When Edwards rode up, he was captured and partially hung again and again from the willows until, half-strangled, he promised to clear out his sheep. In 1899 he sold out and headed for Oregon, where, among gentler people, he became a leading sheep rancher. The new owner, Geddes, tried to graze a flock on the Colorado side and was attacked by masked men who killed nearly a thousand of his sheep. Within a few years some of the attackers went over to sheep themselves as the range grass dwindled to the point of no return. The Little Snake residents turned their attention to driving out the lingering White River Utes, who, although force-marched out of their homeland to a Utah reservation in 1880, came back repeatedly to hunt in their old territory.5 The first settlers had complained that Indians were taking all the game, and even before the battle of Milk Creek and the so-called Meeker Massacre, people said that the Ute reservation had to be opened for settlement—”The Utes must go!”6 Routt County resident Joe Carroll, in a Civil Works Administration interview in the 1930s, gave a firsthand account of one foray against the Indians: I was sixteen years old. . . . The game laws had been in effect for five years and not a single Indian had been arrested. People got to codding the game warden about being afraid of the Indians so he enlisted ten men and set out to arrest 312

him an Indian. The Utes were camped in the bull-berries on lower [Little] Snake River. Nobody was in camp when the warden arrived except the squaws and papooses and a few old crippled men. Everybody was excited and no two men ever told the same story about what happened. Jack White was a broncobuster and fast with a gun. He could lope his horse and pull a gun in each hand and hit two prairie dogs. It’s a cinch Jack wanted to show off. Miles Overholtz was just a big overgrown kid. He shot one of the squaws in the belly. Amos Bennet pumped his gun empty without firing a shot. Al Shaw got hit on the head and took the count when the fight started. When the fight was over his gun was empty. Nobody knows when he pumped it dry.7 This brief account shows the confused attitudes white settlers held for the neighboring Indians with whom they traded, whose tanned hides they bought: the blundering youth, the embarrassed men who emptied their guns or feigned unconsciousness, the show-off gunman, none with obvious sympathy or feeling for the women and children they killed. Many of the cattle ranch headquarters were not in the comparatively lush valley, but in the rugged desert lands. But ranches in and around the Red Desert, even the big ones, lasted only a few years before they changed hands or were seized for back taxes or simply abandoned. For those trying to make a living on the dryland edges, the concept of cash flow was a bitter joke. As time went on, sooner or later nearly all the ranch families retreated to the larger towns—Denver or Chicago if they made a good sale, Rock Springs or Rawlins when sickness, divorce, old age, or the difficulties of remote living in a hostile environment became too much. Ranch families who actually made a modest success of things might move to town and send the children to a proper school. But for most people who lived in or near the desert, extended families, intermarriage with neighbors, familiarity and collusion with outlaws, and a taste for the vast landscapes and

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 313

daily life within the silent reaches made them different from townspeople. Local people were not averse to making a little shady money from investors. One scam was the West Side Mining Company, a fruitless placer gold operation of the 1890s located in the hills southwest of Baggs where the Little Snake leaves Wyoming for Colorado. A local man, John Hardinburgh, sold the site to three investors from Providence, Rhode Island, described by Clarance T. Johnston, their hired surveyor, as “an expert on clam bakes,” a “former milk inspector,” and “a broken down sardine salesman.”8 Horse thieves, suicide, and no gold ended the venture. Surveyor Johnston hit on a characteristic of early Wyoming settlers around the Red Desert rim when he wrote, “For mental alertness the natives of the valley could not be excelled. . . . These [investor] New Englanders, Yankees, . . . were unable to compete with men representing a society stimulated by necessity and purged of the weak and unfit. In every transaction, commercial, social or charitable, the man who lived in the sagebrush walked away with all benefits, prizes and profits.”9 Some of the more interesting residents of the Little Snake Valley were Henry Fraeb, Jim Baker, John Brockmire a.k.a. “Bibleback Brown,” the Reader family, and George and Maggie Baggs. Getting In and Out Some roads and trails that connected the Little Snake River valley to supply centers were old Indian trails, which, in turn, followed ancient game trails, though residents fashioned deliberate shortcuts after the railroad went through. Fresnos, hard labor, and increasing use made these trails into rough roads. As primitive as these important passageways were, they were superior to crosslots travel through the sagebrush.10 The Muddy Trail (so called because it followed Muddy Creek) was good enough for general stage and freight transportation and for moving sheep flocks. Muddy Trail started at Baggs and went fifty miles north to the sheep- and cattle-shipping town of

Wamsutter, which received much of the area’s wool.11 The first stop out of Baggs was Jumbo station. The Muddy Bridge station, near the confluence of Cow Creek and Dry Cow Creek, was better known as Jerry’s Place. A long ride later the Perkins dinner station offered food, drink, and cigars to the jolted, dusty passengers. Next came Willows stage station, described by surveyor Clarance T. Johnston in 1895 as “simply a place.”12 Beyond Willow lay Sulphur Springs station, site of an Indian attack in 1863, on the old Overland Trail where an east-west link connected to the Savery Trail. This connecting link led to Bridger Pass station. Farther north was the Twenty Mile station at Twenty Mile ranch, and then Sixteen Mile station, the last stop before Rawlins. It was a full day’s journey from Baggs to Rawlins or from Savery to Wamsutter in good weather with dry roads, but an overnight trip if conditions were poor. The Savery Creek Trail, which ran from Savery to Wamsutter, linked northwest Colorado and the Little Snake River valley with Rawlins. Several women (mostly hard-pressed widows) drove regular freight schedules between the Little Snake hamlets up to Dad, Wamsutter, and Rawlins in the late nineteenth century.13 Rawlins was a railhead and, in the late nineteenth century, a favorite blowout town as well as a major source of supplies. An important Rawlins store was J. W. Hugus and Company, the first chain store in the Rockies.14 Tom Wise remarked in a CWA interview, “In the early days Hugus & Company did a land office business in Meeker. Some days in summer that main street would be packed solid with four and six horse teams waiting their turn to load at Hugus store.”15 By the close of the nineteenth century the area was dominated by three big ranches—the L7, which mostly ran cattle in the lower Little Snake country; the Leavenworth Cattle Company (better known as the Pothook, for the brand), which dominated the upper Little Snake; and the Ora Haley Company, which ran cows farther to the west in the Bear River area. In practice the cattle from the big three mingled on the open range with the stock of small ranchers, all sorted out in the spring and autumn roundups.

The Little Snake River Valley

313

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 314

Today (2006) the west end of the Little Snake River valley, especially around Baggs, is crowded with gas rig workers and surveyors. The section of the Red Desert recently named Desolation Flats (presumably so people would believe it to be worthless land), fourteen miles west of Baggs, has attracted major energy companies. A pipeline snakes along near remote Adobe Town rim. Motels are full, there are no houses for rent, the trailer park is crammed, the highway is clogged with trucks, local eateries are mobbed, and the methamphetamine problem casts its baleful light over all. Tiny Baggs (population 282 before the coal-bed methane rush) is experiencing a surge of crime. In 2005 the then police chief Mark Lapinskas commented, “The kind of people who come to work here are rough.”16

314

Notes 1. In 1831 an Arkansas-Taos group of would-be trappers, including “3 or 4 Mexicans,” the Sinclair-Nidever party, went north from Taos to the Arkansas River, followed it to its source, then went on to the “Río de los Sozones. . . . Sozone, or Shoshone, being the tribal name for Snakes in New Mexico, doubtless referred to today’s Little Snake River, then known simply as Snake River.” David J. Weber, The Taos Trappers: The Fur Trade in the Far Southwest, 1540–1846 (University of Oklahoma Press, Norman, 1971), 198 –199. 2. Clarence T. Johnston, “The West Side Mining Company,” Annals of Wyoming 7, no. 4 (April 1931), 449. 3. Richard Blatchley and Skylar S. Scott, “A Cultural Resource Survey of the Proposed Sandstone Reservoir, Carbon County, Wyoming” (Office of the Wyoming State Archeologist, 1986), 47, citing Ernest S. Osgood, Day of the Cattleman (1954). 4. John Rolfe Burroughs explains that the Edwards brothers got into the sheep business by ironic accident. Earlier the aggressive Middlesex Cattle Company had advanced on Brown’s Hole with a huge herd of Texas steers. To ward them off, the Edwards brothers converted their own cattle to sheep and made a living sheep fence at the entrance to the Hole. So profitable did they find sheep that they never went back to cows. 5. They had lost Middle and North Parks in 1868 but by treaty held on to the hunting rights to both areas. Few whites accepted their right to hunt there. See Peter R. Decker, “The Utes Must Go!” American Expansion and the Removal of a People (Fulcrum, Golden, Colo., 2004), 111. 6. This rallying cry is still chanted when Colorado football teams play Utah schools. 7. CWA interviews, Civil Works Administration, Moffat County, Colorado, 1933–1934 interviews, pamphlet 356, State Historical Society of Colorado (hereafter “CWA interviews”), 2:217. 8. Johnston, “West Side Mining Company,” 452. Johnston’s account is loaded with humorous sarcasm that spares no one connected with the venture. 9. Ibid., 458. Social Darwinism could fit almost any situation. 10. From Perry Van Dorn, Flickering Lights in Vacant Windows: A History of J. W. Hugus & Co (Pyramid Printing, Grand Junction, Colo., 1992), 24: “Sagebrush country . . . was generally smooth, but around these sage brush was usually a raised mound of earth. . . . [A]s the wheel of the wagon hit this mound of earth, with the brush in the middle of it, the wheel on that side would come to a sudden stop and the forward motion

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 315

of the wagon made the other wheel advance twice as fast. This action caused the tongue to whip from side to side causing a brutal whipping action on the front end of the tongue, being very hard on the wheel team.” 11. In 2006 the paved state highway 789 that connects Baggs and Creston Junction follows the old Muddy Trail for much of the way. As the main connecting passage to the region’s gas fields, the highway bears very heavy truck traffic. Louise Bruning Erb, Ann Bruning Brown, and Gilberta Bruning Hughes, The Bridger Pass Overland Trail 1862 –1869: Through Colorado and Wyoming and Cross Roads at the Rawlins-Baggs Stage Road in Wyoming (ERBGEM, Littleton, Colo., 1989), 78, have Muddy Trail as a Baggs-Rawlins route. It was possible to cut east from the Muddy Trail to the Savery Creek Trail, which did go to Rawlins, but Muddy Trail and Savery Creek Trail were two different entities, though roughly the same as far north as Sulphur Springs. 12. Johnston, “West Side Mining Company,” 453. “There was water and a few willows, also a tent where a half-breed Indian woman served substitutes for food” (ibid.). 13. Helen M. Morgan, Snake River Profiles (Baggs, Wyo., 1970 –1972), 1:7, 27, 34. A local historian names Mag Humphrey, Helen Kelley, Mariah Morgan, and Madeleine Baker Adams as women who freighted and drove trail teams. Madeleine Baker Adams was one of old Jim Baker’s daughters. Her husband was killed in 1896 over water rights, and she was left with four children to support. Mag Humphrey raised nine children, ran a boardinghouse, and, after her husband died, freighted. Helen Kelley and her husband ran stage lines between Slater and Baggs. 14. Hugus got his start as a sutler at Fort Steele. See Van Dorn, Flickering Lights, for a partial history of this enterprise. See also John Rolfe Burroughs, Steamboat in the Rockies (Old Army Press, Fort Collins, Colo., 1974), 40. The only Hugus store still standing and in use in Wyoming is Shively Hardware in Saratoga, once a busy department store stocked with long underwear, coffeepots, yard goods, tack, sundries, and the like. 15. Tom Wise, CWA interviews, 14. 16. Rawlins (Wyo.) Daily Times, December 30, 2003, 1.

The Little Snake River Valley

315

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 316

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 317

23. Red Desert Ranches Annie Proulx

ell before the cattle boom of the 1870s and ’80s, sheep were grazing the Red Desert, and their establishment in Wyoming was part of the larger story of western stock raising, related to the California missions established by Spanish Franciscans.1 In 1824, when Mexico’s constitution declared the nation free and independent from Spain, the missions were farm and ranch centers run by the forced labor of local convert Indians. On the mission ranches horses, sheep, and cattle grew in numbers to the surplus point. During the 1830s the Mexican government turned over choice mission properties to favorites, opening the period of the ranchos. In 1848, California, the great prize of the Mexican-American War, became American territory. That same year gold was discovered on the property of the Swiss-born sheepman Captain John A. Sutter. There was an immediate and enormous rush of people into the territory, so many that the territory became a state a mere two years later. With the gold rush came a demand for meat to feed the miner hordes. Sheep were needed. In the 1850s huge herds came into California from New Mexico, Ohio, Illinois, Missouri, most herds moving over the great trails from eastern Nebraska, through Wyoming, Idaho, and Nevada to California. New Mexican sheep and New Mexican sheepherders moved north into Wyoming. Observant men noticed that sheep driven across the Red

W

Desert did well, finding nutritious forage. In California, as the gold played out and the fever waned, miners turned to sheep and cattle as more dependable sources of income. In the 1860s the movement of the sheep reversed from California to the east during what Edward Norris Wentworth called “the era of the great trails.”2 From northern California the main trail came across central Wyoming on the old California-Oregon-Mormon Trail, but a variant ran south of Salt Lake City into southwest Wyoming and across the Red Desert. The Red Desert was a favored wintering ground to break the journey and recondition sheep before driving them on through the mountains. Unlike cattle, sheep will eat snow for water. Gradually, local men established flocks in and around the desert. The profits were large, the initial investment small. General James S. Brisbin, whose ecstatic 1882 hosanna, The Beef Bonanza; or, How to Get Rich on the Plains, changed so many lives, did not restrict himself to cattle. He described—and named—Wyoming sheepmen as making annual net profits of 35 to 60 percent. The history of the Red Desert is one of early cattle ventures that, after a generation, gave way to sheep raising. There is a common belief today that there were few or no ranches in the Red Desert, that people kept away because of the harsh, arid nature of the place, and its remoteness. Those who 317

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 318

do not know the Red Desert think of it as hot, waterless miles of no-shade sagebrush, but a look at the topographical maps will show numerous flowing wells, many of them rich oases for birds and wildlife, along with intermittent streams and springs. There was water under the desert, and ranchers drilled wells and hauled water. There were scores of ranches, most scattered along stage or freight routes. Others lay in distant and out-of-the way corners. William Logan Jr. (1923–) of the Logan ranch remarked: “There used to be lots of little ranches around the country that had people living on them. So there was a lot more people moving around the area than you would think there was. All those places have gone back to nature. Nobody around them at all. The buildings are all gone.”3 The early cattle ranches in the Little Snake country included the L7, the Pothook, the 71, the Two Bar, Leavenworth Land and Cattle Company, the Figure Four, and many more before Wyoming sheep pressed in. The Logans, too, ran cattle until they sold out in 1947. The desert ranches were generally sheep outfits, as the rough forage and lack of easy water, especially in winter, made cattle grazing problematical. When men with very little money started running sheep and succeeded brilliantly, the rush was on. From the 1880s until after the Korean War the Red Desert was dominated by sheep.4 There were so many sheep on the crowded range that they spilled over into northwest Colorado cow country, precipitating a sheep-cattle war with the state line as the deadline marker. The break came in 1907, when the Forest Service realized that the excess sheep population crowding onto the Sierra Madre Forest Reserve had resulted in severe overgrazing.5 In 1908, following meetings of the Carbon County Woolgrowers Association (sheep interests) and the Snake River Stockgrowers Association (cattle), the area set aside for sheep added Whiskey Park and Big Red Park in the Routt Forest to the permitted range area. Wyoming sheep grazing the forest increased terrifically in number until in 1931 there were 143,274.6 By the 1940s the Forest Service realized that overstocking could not be balanced by man318

agement; cutback programs began. Both cattle and sheep raisers protested, but in 1947, following meetings of the Forest Advisory Board and the Carbon County Woolgrowers, the Forest Service approved a reduction.7 Sheep raising on the desert began to decline in the 1950s in the face of a labor shortage, closure of the railroad shipping points, the traditional three-generation family business run of increasingly fewer assets and more heirs, and more varied occupational choices. Some area ranchers tried cattle. Leonard Hay, of the Rock Springs ranching-banking family, switched from sheep to cattle when he could no longer get New Mexico labor. On the Hay family sheep-raising generations, he remarked: I was the first one in the cattle business. . . . [T]he reason I switched to cattle was that . . . the labor situation in the sheep business was almost impossible. I had straight New Mexico men. Their fathers had come up here and work[ed] and then their sons would come up and work. Many people would not believe what I am going to tell you, but when those kids could get a jalopy for $100, $200, they were more concerned with driving that car than being out in the sheep wagon six, eight, ten months out of the year. . . . And of course then the ski resorts started at . . . [Taos, New Mexico] and they could get jobs there, in the school systems janitoring . . . and in the timber mills and in mines. . . . And the old men got old and retired and the new men just weren’t going to go out and live in a sheep camp. They wanted to drive their car.8 Among the Red Desert ranch names familiar to late nineteenth-century and early twentieth-century Wyoming people were the Albertine ranch, the Barnhart, Red Desert Ranch, the Yancey place, the Hay family’s Middle Ranch, the Daley place, the Karel spread, the Maxon, Willow Creek ranch, Ramsay’s Trout Creek ranch, Cooper’s ranch (east of Ramsay’s ranch near a creek draining southeast from Quaking Aspen Mountain), the McGath place,

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 319

Jarvie ranch, Red Creek ranch, Sage Creek ranch, the Hooten place near Boar’s Tusk, Brooks ranch, the Casa Grande at Black Rock Butte, Bill Lewis’s ranch in the sand hills southwest of Steamboat, Logan’s ranch, the Chilton ranches (including their Steamboat spread thirty miles north of Rock Springs), the Turkey Track, the Antelope, the Jawbone, Bolton’s ranch, Pine Butte ranch, Luman’s, Cow Creek ranch, the Kinney ranch, the Eversole place, and the JO. In the Sweetwater country, ranchers Claytor, Grieve, Beaton, and Sharp (the “Big Four”) ran their stock on the desert. One of the most important early sheep outfits was the Cosgriff brothers’ Cow Creek and Pioneer Sheep Ranch, which ran 125,000 sheep from Fort Steele to Colorado.9 The stories of most of these ranches remain to be written.10 Many passed through dozens of ownership changes, making their history difficult to sort out.11 Almost all of them are now gone, their corrals and bunkhouses returned to the earth. But one or two still stand more or less intact, giving us an idea of the character of early Wyoming ranch life. And some have been studied. The L7 Most historians agree that of the four Swan men who came to eastern Wyoming in the 1870s and established the Swan Land and Cattle Company, three were brothers (Tom, Henry, and Alexander).12 The fourth, William F. Swan, was the son of Henry.13 Alexander Hamilton Swan was the powerhouse brother, although he made most of the business errors of the period and lost the outfit to a consortium of Scots investors in 1883.14 The Scots hired a fellow Scot, John Clay Jr., to run the enterprise. Under Clay’s management the Swan Land and Cattle Company dominated the Wyoming cow business in the early years before and during the boom, when anyone could run cattle on the open range.15 In 1882 William Swan bought the Hat ranch on Pass Creek north of Saratoga. With this ranch, his membership in the Wyoming Stock Growers Association, and his blood connection to the big Swan outfit, locals regarded him as a cattle king. That

same year he moved to Denver, dependent on various foremen and managers to run his ranches, perpetuating the ruinous pattern of absentee ownership so characteristic of the aristocratic style of ranching. In 1883 he started the L7 Cattle Company and the (Little) Snake River Cattle Company, and the next year bought the improvements on a ranch on Cow Creek. This became the 71 ranch, used for hay, winter pasture, and pampering the outfit’s pedigreed bulls. Although the Wyoming newspapers, infused with boosterism, avoided saying anything negative about the terrible winter of 1886 –1887, the spring roundup told the tale.16 Swan’s L7 ranches on the North Platte went into that winter with a “recorded count” of 6,300 cows.17 For weeks the snow was covered with a crust of ice so thick the animals could not break through to grass. The 1887 spring roundup counted only 2,851 head of L7 cattle, a loss of roughly 55 percent. The Carbon County Journal ran a statement by a Saratoga resident: “The roundup is over and the cattle men have disbanded and gone home. From the most reliable sources we learn that the loss has been far greater than anyone anticipated . . . it is fair to say that many men would be happy if they could gather 50 per cent.”18 Swan reorganized the L7 and ordered his surviving North Platte cows down to the protected Little Snake ranch. But if the valley escaped the worst of 1886 –1887, its turn came in 1889 – 1890. Only 300 L7 cows survived at the winter ranch in Colorado, and no more than 1,000 were left on the Wyoming side. A desperate late drive into the northern Red Desert exacerbated the situation. Swan’s foreman, Charley Neiman, described the horror: The range was heavily overstocked with from twelve to fifteen thousand head of stock in each of the three largest outfits—the L7, the Ora Haley ranch, and the Leavenworth Cattle company—and a number of the smaller outfits had about one thousand head each. A number of cattle men decided to drive the stock to the Red Desert country north of Rawlins. But the decision . . . was made . . . too late. It was

Red Desert Ranches

319

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 320

the middle of November before the order to drive was given. Wilson Rankin, foreman of the Haley outfit, was the first to start. . . . I went next with the L7 cattle. . . . By the time I had reached Baggs, some forty-four miles north—I had ten thousand bawling cattle, many weak and in no condition to travel, and many with calves. Drifters and stragglers were numerous. The day before Christmas, somewhere between thirty and forty miles north of Rawlins toward the Red Desert, I turned the cattle loose. They just couldn’t go any further. Many died on the drive. Many were too weak to complete the trip and dropped back. Many died after we had arrived at our destination. It was the worst slaughter I had ever seen. That summer we had branded between 3,500 and four thousand head of calves; the following spring only 174 or 175 calves were branded. . . . The loss of cattle during the winter of 1889 was estimated at 75%!19 After a decade of boom and big beef profits, the bonanza years were over. Mismanagement, absentee ownership, dependence on imaginative “book counts,” the decade of rainy years and mild winters replaced by drought and savage cold seasons, and the overstocking of the range boiled down into a cow pie of economic disaster on the western plains. The L7 folded in 1894. The Pothook, Leavenworth Land and Cattle, and Figure Four Two Idaho cattlemen, Paul Fuhr and W. E. Timberlake, came into the Little Snake Valley in the late ’70s with a herd of cattle. Timberlake immediately sold out to Fuhr and disappeared. Five years later, Fuhr, who had taken the pothook brand, also sold out and went up to Rawlins, where he dealt cards in a saloon. The new owner of the outfit, John H. Durbin, sold the Pothook, as the place was now known, the same day he bought it to one J. B. Insley. That buyer, in turn, transferred the Pothook to the Leavenworth Cattle Company of Leavenworth, Kansas, a company in which Insley was a partner. 320

For the next few years the Pothook was one of the region’s dominant ranches. Mismanagement and the hard winter of 1889 –1890 hurt the company badly. It couldn’t recover and in 1896 went up for public sale. Three Iowa men bought it, using a ranch a few miles west of Baggs as headquarters for the new outfit, with the registered brand of Reverse Four. One of the new owners, Sidney Green, acted as manager, and a nervously energetic local cowboy, Wiff Wilson, headed the roundups. But 1902 – 1903 was another hard winter, and Green took what cash he could get his hands on and headed south. At the 1904 auction the Reverse Four—better known as the Figure Four, since many locals wrote the number backwards anyway—was bought by one E. D. Gould and his silent partner, Ora Haley, for a bargain-basement price.20 Ora Haley, who had a talent for picking out competent men, put local cowboy Wiff Wilson in as manager, and Wilson, promised a quarter interest in the Reverse Four, gave it all he had. The Figure Four prospered under his control until another terrible winter, that of 1908 –1909, wounded but did not kill the Four. Two Bar Ora Haley, of Irish descent, a poor youth from East Corinth, Maine, who ran away from stepmother trouble, came to Denver in 1865 as a bullwhacker, one of the classic routes into the west. He started his adult life in Laramie, running a butcher shop that had a contract to supply meat—any meat—to the Union Pacific Railroad, then under construction. The shop sometimes ran low on venison or elk. Eventually, to maintain a steady supply, Haley acquired a ranch on the Laramie plains. Haley was a shrewd businessman who knew cows from their anatomical structure to their daily market value. He kept good books. He lost most of his early herd in the winter of 1871–1872 when the cows fell through frozen Lake Hutton southwest of Laramie. That winter forced many area cowmen to the wall, and the local banks would not loan Haley money to start up again. He

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 321

showed his books to New York bankers and got a line of credit that put him back in the cow business. A few years later he pulled off a coup. The shorthorns the Oregon emigrants had driven across the country decades earlier had multiplied tremendously in the Pacific northwest, and that range was overstocked. Haley and a few hands went to Oregon, where he bought a large herd for $5 each, drove them to Omaha, where he sold them for $50 dollars a head. In later years Haley made quick profits on short-term market tides. In 1894 he bought the old Majors-Saintsbury ranch on the Little Snake River. Local cattlemen saw their new neighbor and his Two Bar brand both as an intrusive threat and as a sugar pot from which they might enlarge their own herds. Haley’s foreman, Hi Bernard, was an exceptionally astute cowman, and it was said he made Haley a million dollars. Haley enraged Routt County residents when he put an ad in the Craig Courier in July 1896, canceling the western tradition of free meals and bedroll transport for representatives from other ranchers during roundups.21 At that time Brown’s Hole in Colorado was dominated by the so-called Bassett Gang, a group of local ranchers headed by the narcissistic “Queen” Anne Bassett and Madison “Matt” Rash, a Texas trail driver who had come north in 1882. Rash and Bassett were believed to be cattle rustlers, especially of Ora Haley Two Bar stock. The pair saw Haley as their major enemy and made him the most hated man in the region. Bassett, a complex and handsome woman of strong character, drove many of Haley’s cows into the Green River, where the current carried them away. Matt Rash was found murdered in his cabin in 1900 after ignoring a note on his door telling him to clear out. At this time the notorious Tom Horn was in the area and had even worked for Matt Rash under the alias “James Hicks.”22 In 1904, in one of those extraordinary twists that characterized Brown’s Hole events, Hi Bernard, Ora Haley’s right-hand man, married Anne Bassett, twenty years his junior and his employer’s arch enemy. When Ora Haley heard the news, he fired Hi Bernard. The angry and empty marriage lasted until 1911, when Bernard and Bassett divorced.

In 1910 Ora Haley’s new manager, Bill Patton, threatened harm to Harry Ratliff, the U.S. Forest Service acting supervisor of the newly created Park Range Forest Reserve, if he tried to keep the Two Bar cows off forest land where they had grazed for years. In the Little Snake River valley a group of local men, including Wiff Wilson, Bill Patton, Al Reader, and others, plotted against Ratliff. Ratliff later wrote: “[I]t was decided that my activities were not in accordance with the demands or policies of the Class C units, and that steps should be taken to get rid of me.”23 The cattlemen worked up a complicated plan that led to Ratliff’s arrest for fraud in a horse deal made some time before he joined the Forest Service. Out on bail, and with trial set for September, Ratliff conducted a big roundup of cows grazing on national forest land and found “something like 46,000 head of cattle on range supposed to be pasturing 20,000 head.”24 Many of those cows belonged to Haley, who had no permit. After two trials Ratliff was acquitted, but the hatred local cowmen felt for the Forest Service and its rangers persisted for decades. In 1914 Ora Haley sold out his cattle interests in northwest Colorado. The Figure Four, still in operation under different management, was now the single big cattle company remaining in the Little Snake River area. The end of World War I knocked the demand for beef for a loop, and the 1916 Homestead Act brought in a flood of settlers. The range was laced tighter than a corset with barbwire and new people. The day of the open-range-and-roundup cattle business was nearing sunset. Sheep claimed the Red Desert. Circle K Irishman Tim Kinney began his Wyoming adventure as a construction laborer on the Union Pacific Railroad, Carbon Division. Kinney was short, muscular, and ambitious and worked his way up. When the track reached Rawlins, he became the first station agent there and later worked as agent in Rock Springs. Kinney eventually struck out on his own. He bought some cattle, registered his Circle K brand, and ran the stock in the Bitter Creek re-

Red Desert Ranches

321

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 322

gion of the Red Desert. Writes John Rolfe Burroughs, “It wasn’t long until his ‘Circle K’ outfit controlled an immense piece of country.”25 While most cattlemen thought the area was all sage and greasewood, Kinney, who undoubtedly had watched the trailed sheep herds come through, recognized that there was plenty of nutrition on the thousands of square miles of presumably useless land, and what was good for sheep might be for cattle as well. Half a century later R. H. Burns noted in Wyoming’s Pioneer Ranches: “[T ]here is the Red Desert, once considered worthless but now one of the finest winter livestock ranges in the west, thanks to the low growing salt sage whose low clumps provide a feed equal in food value to high quality alfalfa hay. And this apparently worthless barren land has untold and unfound riches underneath, in the form of precious and semi-precious minerals.”26 As Kinney increased his wealth and power, he got into storekeeping and banking in Rock Springs. He became a leading businessman of the day. Kinney was one of the original stockholders in the famous and still-powerful Rock Springs Grazing Association. This stock growers’ association later had connections to every imaginable business activity, from sheep and cattle to railroads, gas and oil, hunter outfitting, groceries. Kinney was doing very well but, says Burroughs, ran into trouble in the form of the intrusive Middlesex Land and Cattle Company, whose manager’s name, Fred Fisher, continued for many years to represent tyranny in southwestern Wyoming. Kinney was not entirely wiped out by Middlesex—the winter of 1886 –1887 finished him. Undaunted, and with excellent credit, he switched to sheep and built up a flock of 50,000 ewes, which he pastured on the Red Desert, setting the stage for a century of ovine grazing. He was an important figure in the historical shift from cattle to sheep. Kinney was also one of the 1905 founders of the Wyoming Wool Growers Association. The organization started at the same time as many government organizations and programs, including the U.S. Forest Service, and had dealings with the federal government from the start. 322

Today, out of the once-many small outfits in the Red Desert, two nearly complete ranches — the Cow Creek ranch, south of Adobe Town, and the JO, east of Dad—and fragments of a third, the Jawbone, near Buck Draw, still stand in the desert. They are irreplaceable historic remnants of Wyoming’s legendary ranch past. Cow Creek Ranch The Cow Creek ranch, south of Adobe Town, persists. Like most sheep outfits, it has a row of many buildings, some old log cabins, some stick-built. Of interest is an old log barn with a smithy and a small log cabin near the ranch house. Reputedly the ranch once sported a sign reading, “Sixty miles from town, thirty miles from water, two feet from hell.” A tattered Confederate flag flew for some years until it disintegrated. Poachers have always been thick in the Red Desert, and, according to Ron Lockwood, a biologist with Wyoming Game and Fish, one cabin on the Cow Creek ranch was set up as a court where a judge took up temporary residence. Rangers and wardens brought miscreants to the instant reckoning. The original Cow Creek and Pioneer Sheep Company ranch was not owned outright by the famous Cosgriff Sheep Company, one of the major Red Desert graziers, but it was linked to them through one of James E. Cosgriff’s many partnerships. John K. Hartt of Rawlins bought Cow Creek land and used the name “Cow Creek Ranch.” It was one of the region’s biggest outfits. A story circulated that Hartt, as a young man traveling through the west, had been forced to leave the train at Rawlins when his money ran out. He made a fortune with the Cow Creek and another outfit, Yellowstone Sheep, operating out of Lander. His Red Desert winter range included the Hiawatha, Sand Creek, and Powder Springs areas. Summers, Cow Creek sheep ranged along the Little Snake tributaries. At the turn of the century the Cow Creek ranch was involved in the illegal sheep push into Colorado.27 Later, in 1909, at the end of the Wyoming-Colorado sheep-cattle war, Wyoming sheep were trailed from the Routt forest to Steamboat for shipping with-

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 323

out trouble despite hot talk, although a man who had once worked for the Cow Creek ranch, “undergoing an attack of bad whiskey,” committed suicide in Steamboat at the time and a newspaper hinted he had been killed by cattlemen.28 Eva McCormick Corson, eighty-six when she was interviewed in 1989, came to the Cow Creek ranch in May of 1917, when she was fourteen. Her stepfather, Roy Eversole, worked for the ranch. She remembered that “when I first got there they were building a new [headquarters ranch] about ten miles south of where their old headquarters was. The old one they had was nothing but a little log cabin and a few little buildings or two. . . . It was all just a sheep headquarters. Had all the feed down there and then they had all the supplies for the sheep camps.”29 Her chores included caring for stock, cooking for the men, and general household and ranch work: We always had horses and things to take care of, [men] to cook for and such as that. . . . There was truck drivers and then they had some freight outfits, horses and sheep herders. . . . [I]n the summer-time there wasn’t anyone there but just us, cause they went up into the mountains for the sheep and summered ’em in the mountains. Then they came back down in the fall. They had to leave the Forest reserve in September and come back down. They was there all winter and brought their sheep down. They had about . . . I would say, fifteen or more herds [bands] of sheep and they had all the sheep herders and that and they had to come in there for supplies, to the ranch.30 Her mother and stepfather homesteaded in 1919 on land once part of John Hartt’s holdings: “A fellah by the name of . . . ‘Readystep,’ . . . hadn’t ever homesteaded. He just took a squatter’s right there. And then he got pretty old . . . and he had a sister in California so he went out there . . . and that’s when we got the place. My dad set then and he filed on it and homesteaded it and that’s how we got it.”31

In 1923 Eva McCormick married John Corson of the Antelope ranch, a few miles to the north. Later the couple lived at the Cow Creek ranch. Eva Corson’s stepfather, Roy Eversole, worked on the Cow Creek Reservoir. “Then in 1952 it [the ranch] was washed out, gave out on us,” Mrs. Corson recalled. “We was living there then, my husband and I was living down on the [Cow Creek] ranch then and the folks was up at their place and it went out and it wasn’t long after that that the [Cow Creek] outfit went out of business. They just parceled it out.” Mrs. Corson thought family dissension was the downfall of Hartt’s Cow Creek Company:

Well, the main thing of it, the owner of it died, old John Hart is his name. He had his wife and six kids, six girls and I don’t know how it come about but in some way they couldn’t agree in some ways and they just finally wound up in some kind of a mix-up and so they give it up and so they sold it just in parcels. They sold all the sheep and they sold the land out in parcels and I don’t know who’s got it [Needmore, the parents’ homestead] now. . . . It run for a long time. It was an old, old outfit but after the old man died then it just seemed like one thing and the other went wrong. Apparently after the Cow Creek Company breakup, the Corsons acquired the ranch. During World War II the Corsons went to Cheyenne and worked at war jobs. After the war they returned to the ranch until John Corson’s health began to fail: “[W]e had to leave cause he wasn’t able to work. He had been sick for quite a while and dad and my mother they was get’n old and neither one of em’s able to do anything so that’s when Elzy [Eversole] got the [parents’] ranch. So we left there and moved in here [Rock Springs] and been here ever since.”32 Today the Cow Creek ranch is slowly disintegrating, the buildings in need of repair, a curiosity to the many energy extraction

Red Desert Ranches

323

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 324

crews driving past. The Bureau of Land Management uses the yard beside the old barn as a staging point in annual wild horse roundups. During a roundup it is common to see helicopters parked beside the ancient log barn, well-worn saddles and feed stored inside. JO Ranch Of the many ranches that once dotted the Red Desert only the remote JO exists as a historically significant representative of the early sheep business in Wyoming. It lies on the northwest side of Cow Creek, east of Dad, in an area of intense gas exploration and extraction. The complex exists as fifteen buildings arranged much as those at the Cow Creek ranch are—in a long row. Several buildings are handsome structures of dressed stone probably dating back to the 1890s, once covered with an adobe-like coating decorated with incised lines, although they, as the other buildings, need repair and maintenance. The original appearance can be seen on the lee side of the buildings. The ranch is considered eligible for listing with the National Register of Historic Places as representative of Wyoming’s pioneer sheep industry. With repairs and maintenance it could become an important historical exhibit of an early Red Desert ranch. Joseph Rankin, the Rawlins livery stable owner who guided Major Thornburgh’s march on the White River Ute reservation during the 1879 so-called Meeker Massacre and who made a famous ride to bring help to the beleaguered troops, took patents on adjoining parcels of land in 1885, 1888, and 1890, the beginnings of the JO.33 His brother, James Rankin, who was a partner in the livery stable, also took up land in the area. In 1899 Joseph Rankin sold the JO to the Carbon County Sheep and Cattle Company (Dow and John Doty), and four years later sold the livery business and moved to Utah.34 By the 1920s and ’30s the Carbon County Sheep and Cattle Company was dominated by the Tierney family, sheep ranching partners with William Niland and M. Glenn. During this period 324

the shearing pens at Dad burned, and as many as 10,000 sheep were sheared annually at the ranch, using itinerant shearers from New Mexico and Texas. Summer range was near the Hahn’s Peak–Columbine area in Colorado. In 1953 Carbon County Sheep and Cattle sold the ranch to a Little Snake sheepman, George Salisbury, descendant of early settlers. In 1964 Salisbury sold to a consortium of local ranchers, the Eureka Pool, who installed livestock scales, corrals, and chutes and ran cattle instead of sheep. Eureka sold to Tom Grieve in the 1970s, who ran both cows and sheep until the early 1990s, when he sold to Pittsburgh and Midway Coal Mining Company, but continued to lease and run the ranch. In 2005 the BLM acquired the ranch from Pittsburgh and Midway in a land swap. The log ranch house (its logs hollowed with woodpecker holes) is the first building encountered after entering the gate and is believed to once have served as the Willows stage station on the Rawlins freight road north of Baggs. After the stages closed down, Red Desert ranchers helped themselves to the buildings. Buildings of particular interest are the handsome stone dining hall/kitchen and the stone bunkhouse. Ranch historians will find the intact blacksmith shop, its rafters decorated with hundreds of coyote paws, the log, stone-floored barn, the shearing shed and addition where the wool was tromped into threehundred-pound sacked bales, of great interest. The walls and beams of several of these buildings are covered with the inscribed and painted names and sometimes addresses of shearers and trompers. Jawbone Ranch The large barns of the Jawbone ranch loom on the crest of a rise near the old Buck Draw shearing pens west of Rawlins. A recent survey indicates that the buildings are in poor condition, and “[d]isturbances to the ranch area include erosion, livestock use, decay, and probable artifact collection.”35 Its major attraction is to historians interested in Wyoming ranches, for it lies in the heart

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 325

of important sheep country associated with the Cosgriff, Savage, Miller, Hadsell, Riner, Daley, and Stratton outfits. The Jawbone came into existence in the 1960s when A. D. “Doc” Fulton bought the Sundin-Johnson Sheep Company after working as foreman for the outfit for some years. In partnership with Tom Stratton of the Sage Creek Sheep Company, the SundinJohnson was renamed the Jawbone because, said John Niland, “according to them it was financed with talk.”36

Notes 1. Richard Blatchky and Skylar S. Scott, “A Cultural Resource Survey of the Proposed Sandstone Reservoir, Carbon County, Wyoming” (Office of the Wyoming State Geologist, 1986), 47. The authors also note that Jim Bridger kept “a small flock [of sheep] at Fort Bridger as early as 1845.” 2. Edward Norris Wentworth, America’s Sheep Trails (Iowa State College Press, Ames, 1948), 258. “The days of the great trails break into three periods—1865 to 1880, when the bulk of the drives were breeding sheep; 1880 to 1885, a period of transition; and 1885 to 1901, when the great majority were wethers” (ibid.). 3. William B. Logan Jr. interview by Shirley Black, n.d., Western Wyoming Community College Oral History Project (hereafter “wwcc Oral History Project”), 1:14. 4. John Niland, A History of Sheep Raising in the Great Divide Basin of Wyoming (Lagumo Corp., Cheyenne, 1994), 48 –59. U.S. Forest Service, History of Routt National Forest 1905 –1975 (1965; revised in part, 1975), sec. 6, “Grazing,” 47–56 (hereafter “usfs Routt”). 5. From usfs Routt, 48: “It is recorded that the Sierra Madre . . . issued permits on a prior use basis for 300,000 sheep in 1906 on range that in 1935 grazed 45,000 sheep and still was overused.” 6. Ibid., 50. “When the door opened, the excess sheep from the Medicine Bow literally poured into the Routt. The flood was so intense that in 1915 the Supervisor reported: ‘By 1916 there will be no unused range on the Routt’” (ibid.). At first the sheep were permitted only on the higher elevations, but in 1923 the Forest Service began eliminating lowerelevation cattle range. Once eliminated from usfs, the land was homesteaded. When the homesteaders failed to make a living in the 1920s and ’30s, sheepmen bought them out. In some cases homesteaders were helped along by regional hostility. 7. From usfs Routt, 54–55: “This action doubtless was a major factor in precipitating the investigation of the Forest Service range management policies by the Barrett Congressional Subcommittee on Public Lands. . . . [T]he history of grazing on the Routt has been in three phases: (1) An invitational phase to use virgin range which lasted from 1906 until about 1915; (2) a phase of bewildered realization that all was not well . . .; (3) a full realization that the range resource is tied in with the soil and water resources and that conservative grazing must be reached quickly to prevent utter destruction.” Niland, History of Sheep Raising, listed fortyfour sheep brands in his invaluable memoir.

Red Desert Ranches

325

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 326

8. Leonard Hay interview, 1989, wwcc Oral History Project, 7. Basque sheepherders dominated the industry in the northern Big Horn range. Red Desert sheepmen preferred Mexican men from New Mexico and Texas. 9. Wentworth, America’s Sheep Trails, 318 –319. 10. Sources of some information on Red Desert ranches include Ruth Beebe, Reminiscing along the Sweetwater (Johnson Publishing, Boulder, Colo., 1973); Jack Price, Wild Horse Country in Wyoming (Pine Hill Press, Freeman, S.Dak., 1996); and Niland, History of Sheep Raising. The Wyoming State Historic Preservation Office has a great deal of useful information on old ranches. Mrs. Beebe states that many ranchers in central Wyoming came originally from County Cork, Ireland, especially from small towns “on Banty Bay in south Cork.” She likely meant Bantry Bay on the southwest coast. Niland details which shearing pens each Red Desert sheep company used: Walcott, Bolton, Rawlins, “Poop Lagoon,” Daley, Wamsutter’s “Irish” pens and the Stratton “English” pens, Snake River, Copenhagen, Red Desert, Siberia Ridge, Hadsell, Mahoney, Buck Draw, and Four Mile. Eva Corson, who lived on Cow Creek ranch from 1917 until the 1950s, remembered the Bitter Creek shearing pens. wwcc Oral History Project interviews with Rose Teters Logan, William B. Logan Jr., Eva Corson, and others give details of desert ranch work in the early twentieth century. 11. For example, William Logan describes the fortunes of the family ranch after it was sold in 1947: “The ranch was originally sold to George Froggett and Bill Logan had the Toastmaster Ranch and they bought the Goettche ranches and bought our ranch and then it was sold out to a fellow by the name of Frank Bosler, Bosler sold it to Leonard Hay, Leonard Hay sold it to somebody else. I think it changed hands about six times. Each time became a little more run down and no use on it or anything and the price went up a little bit each time.” William B. Logan interview, wwcc Oral History Project, 9. 12. There were reportedly eight brothers, the sons of a Scots-Welsh farmer who had settled in Greene County, Pennsylvania. 13. Lawrence M. Woods, John Clay, Jr. (Arthur H. Clark, 2001); Maurice Frink, Cow Country Cavalcade (Wyoming Stock Growers Association, Cheyenne, 1954). 14. Alexander Swan left Pennsylvania, tried Iowa. He later told Wyoming historian C. G. Coutant, “It was too tame for my ambitious nature and I longed for the more perfect freedom of a new country. My ambition was to rear and handle cattle, but the fond dream of my life was never realized until I reached the great grazing regions of Wyoming. There was enough room in Wyoming to satisfy my fondest ambition and 326

hopes.” C. G. Coutant unpublished articles, cited in Paul M. Edwards, “The Scottish Role in Midlands America with Particular Reference to Wyoming 1865–1895” (Ph.D. diss., University of St Andrews, 1971), 142. 15. John Clay Jr.’s My Life on the Range (1924; reprint, Antiquarian Press, New York, 1961) is a highly regarded authority on the cattle business from the early 1880s to the 1920s. His book is by turns sentimental, hardheaded, and snobbish. He respected successful businessmen. He disliked cowboys, nesters, wire, and especially did he despise Alex Swan, whom he characterized as vain, jealous, overbearing, surrounded by sycophants, and—Clay’s most derogatory epithet—a poor businessman. 16. See the chapter titled “The Winter of Death” in Helen Huntington Smith’s The War on Powder River (University of Nebraska Press, 1966, 1967) for a graphic description of the season and a discussion of the Wyoming newspaper editors’ dodging of bad-weather news. 17. A recorded count was an actual count, in contradistinction to the infamous “book count” of wild guesses and nonexistent cows common in the boom period of western cattle raising. 18. Herbert O. Brayer, “The L7 Ranches,” Annals of Wyoming 14, no. 1 (January 1943), 22. 19. Charles Wilson Neiman, cited in Brayer, “L7 Ranches,” 26. 20. John Rolfe Burroughs, Where the Old West Stayed Young (Bonanza Books, 1962), 261. 21. cwa interviews, 161. Signed by Haley, it read: “From and after July 1, 1896, all persons coming to our round-up wagons and ranches for accommodations, will be charged 25$ per meal and 25$ per day for wrangling horses and hauling bedding. It has been the custom of certain parties to take up and ride horses belonging to our outfit and also to drive our bulls off the range to be run with their cattle. Hereafter in no case will any such person be allowed such privileges.” 22. Burroughs, Where the Old West Stayed Young, 206. 23. Ibid., 286. Class C units were big cattle outfits. 24. Ibid. Burroughs also relates the story of a confrontation between the rangers and Haley’s men who were driving Two Bar cows into Slater Park. Haley’s lawyer got the Two Bar a permit to graze the usual number of cows on the forest land, an outcome that further enraged Routt County people. 25. Ibid., 59. 26. Robert Homer Burns, Wyoming’s Pioneer Ranches (Top-of-theWorld Press, Laramie, 1955), 561. “Salt sage” can refer to several species of Atriplex. The group includes the fourwing saltbush, shadscale saltbush, and Gardner saltbush, all excellent browse for antelope, deer,

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 327

sheep, and cattle. These plants are adapted to alkaline or saline soils. Winter Guide to Native Shrubs of the Central Rocky Mountains, ed. William T. McKean (Department of Game and Fish, State of Colorado, 1956), 57– 61. Dennis H. Knight comments in his Mountains and Plains (Yale University Press, 1994), 115: “Saltbush desert shrubland is perhaps the most arid vegetation type of the intermountain basins. It is characterized by sparse plant cover and the prevalence of Gardner saltbush, a short, evergreen shrub also known as saltsage. . . . Saltbush desert shrubland occurs commonly on fine-textured soils developed from shale or alluvium.” 27. usfs Routt, 47: “Sheep owned by Cow Creek Sheep Company were caught in Red Park in 1903 and were run out by 300 riders in one of the greatest sheep stampedes in history. No one was killed.” 28. Ibid., 49. 29. Eva Corson interview, 1989, wwcc Oral History Project, 2. All quotations of Eva Corson come from the same interview. It is not always clear which ranch Mrs. Corson meant in her remarks. 30. Part of Red Desert ranch life at that time included the famous Thanksgiving parties and an annual picnic at Pine Butte Ranch. Remarked Mrs. Corson: “Well that snow would lay in that timber up there ’til purt’ near every Fourth of July why they’d be a bunch of us go up there and make ice cream and all of that.” William Logan Jr., wwcc Oral History Project, 11–12, gives a vivid account of the Fourth of July picnics on Little Mountain, where the snow lasted longer than on Pine Mountain. 31. They called it the Needmore Ranch. Today it is the Eversole ranch, run by John Eversole, the son of Elza Eversole. 32. Elza Eversole was a well-known cowboy and rancher in and around the Red Desert and Little Snake Valley region. He was also a wild-horse catcher in the 1930s. He acquired not only Mrs. Corson’s parents’ Needmore Ranch but also, after John Corson died, the Cow Creek ranch and its mineral rights, which he shared with Mrs. Corson. 33. See the section on Fort Steele in Chapter 18, this volume, for a discussion of the Milk Creek fight. 34. Much of this information on the JO ranch is drawn from Robert G. Rosenberg, “Report of Historical Investigations of the JO Ranch (Site 48CR1203), Carbon County, Wyoming” (prepared for the blm Rawlins Field Office, 2005). 35. “Class III Inventory, Jawbone 33-1 Connector Road” (trc Mariah Associates Inc., for the blm Rawlins Field Office, 2005), 1. 36. Niland, History of Sheep Raising, 211.

Red Desert Ranches

327

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 328

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 329

24. Horse Bands of the Red Desert Annie Proulx

n 1519 Hernán Cortés brought sixteen horses to the New World — eleven stallions, including two pintos, and five mares. Most of these horses were the Spanish Cordoba strain with Barb antecedents. Although it is also a hard-hoofed desert horse, swift and hardy, like the Arabian, it is different, with a slightly convex head profile and a low-set tail. Cortés himself rode the Black One (El Morzillo), which died. The remaining horses became the foundation herd of North America. As the Spanish established missions and explored the southwest, at first they tried to keep the horse from the Indians, and perhaps the tribal sport and avocation of horse stealing started in defiance of their prohibition. The mustangs multiplied fantastically on the southwest range, “corresponding in climate and soil to the arid lands of Spain, northern Africa and Arabia in which they originated.”1 At the California Spanish missions, horses proliferated so rapidly their numbers interfered with cattle and sheep raising, and they “were frequently rounded up so that the excess could be killed by spear thrusts.”2 By 1900 their descendants numbered more than a million mustangs.3 Many people, when they hear the words “Red Desert,” think of the free-running horses that live there, believing they are descended from the stock loosed on the continent by the Spanish

I

conquistadores and perpetuated in Indian herds. A dark dorsal stripe and striped legs and hooves are held up as proof of their Spanish ancestry. Others say the horses are feral— domesticated animals that have escaped from ranches or have been deliberately turned loose onto the desert, where they revert to the ancestral ways of horses. Genetic testing has helped to clarify our views of the mixed nature of desert horse ancestry. And, no matter what their descent lines, the proliferation of horses on the Red Desert has had an impact on wildlife. The mustang stock absorbed the horses of settlers, exploration expeditions, the army, emigrants on the trails, horse herds driven east from California and north from Texas; gradually they watered down (or improved, depending on your viewpoint) the mustang stock of Spanish origin. The mustangs were small horses, thirteen to fourteen hands high, and Americans have always believed bigger was better.4 Smallness was seen not as an adaptation to habitat but as degeneration. The short-backed mustangs had five lumbar vertebrae rather than six, which let them carry disproportionately heavy loads. Don Worcester wrote, in requiem, “By the 1890s the game little wild horses were nearly gone from the southern ranges and were fast disappearing from the northern plains.”5 Dobie, writing in 1934, said bitterly, “There are still a few wild horses in Nevada, Wyoming and other Western states. Only 329

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 330

a trace of Spanish blood is left in most of them. When caught they are canned into dog food.”6 Even today some Red Desert animals are genetically similar to the old Spanish horses from the sixteenth century and later Indian herds, as genetic tests after the annual wild horse roundups showed in 2001 and 2003. The Lost Creek and Stewart Creek herds, and some Red Desert herds north and south of the I-80 highway barrier, show high percentages of Spanish characteristics, according to Chuck Reed, a wild horse specialist for the Bureau of Land Management. The relatively isolated Lost Creek band is, he says, “a little Iberian remnant.”7 A campfire conversation that Jack Price had with Frank Robbins, the supreme Red Desert horse-catcher in the 1930s and ’40s, is interesting. Robbins was speaking about the Chain Lake horse bands: It was an area alive with wild horses of a strain not common to the rest of the desert. . . . [Frank’s] favorite horse, Old Buck, was a Chain Laker, as well as Parkoon and Chipmunk. These horses were roans, standing up well on long legs. Old Buck showed buckskin breeding, having the dorsal stripe and the jackass stripes behind the knees. In winter his coat was buckskin color, while in summer after he had shed, his color was roan. Robbins said when he was in the Cavalry [World War I] breaking horses for the army, he saw many horses shipped in to be broken. Horses just like the Chain Lakers came from Texas. They were called Copper Bottom strain. They broke out well and were good horses like the Chain Lakers were. He said he felt certain the horses were of the same breeding. The original stock probably being of Spanish origin.8 The first ranchers around the edges of the Red Desert turned their stock out onto the unfenced desert to take care of itself. John F. Gooldy, son of an 1888 settler in the Little Snake Valley, wrote: “[M]any of the settlers did not think they would ever need to feed 330

their cattle and horses, and made little attempt to raise feed. Cattle were wintered on the range from west of Baggs to Green River. The calves went to the desert with the cows. Horses could winter most anywhere, and did.”9 William Logan Jr., a neighbor of Jack Price’s, remarked, Now the horses that you used, they were all the wild mustangs. . . . A lot of them came from the Firehole area and some came from what is now known as the Marsh Creeks over along Flaming Gorge reservoir. They made better saddle horses than the ones from the Firehole area because they were descendents of some of the freight teams that used to come through. . . . Some of the people that ranched in the Manila area had pretty well bred horses over there. Hamiltonian [sic] and what they call steeldust horses and they got loose into those areas and they are descendents of some of those pretty good horses. They made excellent saddle horses.10 Milford “Toots” Montgomery (b. 1891) of the Standard ranch above Savery trapped horses in the Red Desert most of his adult life. A profile of him, written in the 1970s, notes, “Toots ran . . . his wild horses on the desert north and northeast of Baggs. His wild horse traps are still standing and some still in use. He has probably trapped 1,000 head of horses, which he has shipped as far away as Los Angeles, California, Clovis, New Mexico, and Fort Smith, Arkansas. . . . He has put good blooded stallions with the wild ones, so in the past few years, his wild horses were not merely mustangs but good quality horseflesh, with the vitality and stamina of the wild horse and the build and horse sense of the well-bred western horse.”11 Horse-drawn freight lines persisted into the twentieth century, and some Red Desert ranchers, as Jepp Peterson and Jim Hansen, raised Percherons and other draft horse breeds for this market. Along the northern edge, between Steamboat Mountain and Chalk Buttes, Jake Price and his five sons caught and sold wild

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 331

horses. Jim and Doyle Saglauska concentrated on the area then called the Marsh Creeks or the Marshes, which lay between Little Mountain and the Green River, about thirty-five miles south of Rock Springs. Bill Lewis, “Steamboat Bill,” who ranched in the Sand Hills at the southwest end of Steamboat Mountain, and, on the Blue Rim, Elza Eversole, Harold Anderson, Tom Pulley, and others ran horses in the ’30s and ’40s. In the Washakie basin a loner, Tex Love, was a horse catcher.12 Two railroad towns, Wamsutter and Point of Rocks, had pens and stockyards where caught horses could be held until they took the fatal train ride. Although horse catching was considered exciting fun, there were perils. Bill Logan Jr. described the death of Alec Logan: [H]e was chasing wild horses at that time and he was over toward what is the old Currant Creek ranch. . . . He was quite a cowboy and was riding a real outlaw type horse and the horse was bucking toward some old barbed wire fence with him and he didn’t want to hit that fence and be tangled up and cut up, so he baled [sic] off the horse and was kicked in the leg as he went off. At that time, there were telephones at the ranch. . . . [M]y dad got him sort of situated, . . . rode over to the Currant Creek ranch and telephoned in. . . . [Alec] had crawled along toward where the road was and they picked him up and took him to town and he was in very good spirits all the way. Pretty tough kind of fellow. Joked and laughed with them coming to town but later on that night they came back out to the ranch, had to be somewhere in the early morning hours and reported to us that he had died of a blood clot.13 Cowboying in the Red Desert was tough and full of discomforts. “Doc” Chivington commented dryly on an interrupted spring shove onto the desert: Gilbert Wilson [a son of Wiff Wilson] and myself packed our outfit and came to Powder Springs to gather saddle horses

that had crossed to the west side of Snake River during the spring while the water was low. We rode several days, found the horses, and were getting ready to start the spring shove up when [Wiff ] Wilson sent Leanord [sic] Bean out to tell us to meet another herd of steers in Wamsutter. Gilbert had captured a wild horse so was leading it. He started to the home ranch while Bean and I packed our outfit and headed for Wamsutter. I had been on the desert so long drinking any and all kinds of bad water that I was beginning to feel the effects of it badly. When we came by South Barrel Springs [earlier,] there was a large snow drift there which we melted and used the water from it. On the way back I had a vision of that snow drift and was looking forward to it still being there, but when we camped there my snow drift had all melted. We had to use the rotten alkali water. I had a raging fever that night and slept very little. . . . I went to Rawlins and saw the doctor. He gave me a bottle of Sal Hepatica which I was supposed to take, but never did. When a man had seven or eight hundred steers on his hands and short handed besides, he did not have much time for taking medicine. What I needed most was a square meal without baking powder bread which ruined more men’s stomachs than it ever helped and I also needed a few good drinks of water.14 The Red Desert served as an immense horse pantry for Wyoming people, and when the national economy slumped into the Great Depression of the 1930s, the horses of the Red Desert saved many ranchers from outright destitution. From the 1920s onward the increasingly urban American population, removed from the farm, where offal, orts, and bone scraps went to the family cats and dogs, became dependent on canned pet foods. In those cans was horsemeat, the meat of the wild horses of the western states rounded up by money-strapped ranchers and cowboys. As tractors and cars replaced draft and riding horses in the 1920s,

Horse Bands of the Red Desert

331

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 332

there was a sudden horse surplus and a markedly different way of seeing the horse, not as a useful helper, but as a hungry animal that took much trouble to keep. Horses were turned off farms and some ranches to look after themselves as the pet food market was growing.15 Chicken farms, too, especially in Petaluma, Hayward, and Vernon, California, wanted horsemeat for feed and began to buy up horses around 1920. The railroads, always eager to serve freight customers, cooperated with a special chicken feed shipping rate.16 Fox farms and mink farms bought Wyoming horses.17 The Red Desert supported hundreds of horse catchers. So intense was the trade that E. H. Edwards says 100,000 western horses were killed each year. By the 1970s the number of horses on the western range had dropped to about 18,000 scattered head. Not all of the feral horses went to dog food cans. Range horses were often broken for ranch use. John Rolfe Burroughs says, “A few, notably the Bar Ell Seven, rounded up or bought wild horses, which were plentiful in Southwestern Wyoming and Eastern Utah, and broke them out for use as cow horses.”18 Remarks John Rodney Barney in his Looking Back, an account of the hardscrabble Wyoming cowboy life, “If we could catch the wild-horses and break them before they got too old, they would make awful good tough horses to ride. They could go all day and not get tired. . . . Those horses were good rustlers and they were able to forage real good for food and water. . . . Most of all the wild-horses had good hard feet. A lot of them had black hooves which seemed to hold up better than the white ones.”19 Horse catching was part of the Red Desert local economy, and it was rural entertainment. Even miners from the Rock Springs, Superior, and Hanna coal mines in the northern half of the desert trapped horses when they could. William Logan Jr. remarked, “There were a lot of people from Rock Springs that loved to run wild horses and that was a great sport. Lot of thrills and spills and chills, chasing wild horses. So these fellows would come out and help do the ranch work and then everybody would take off and spend two or three days just chasing wild horses . . . had a lot of fun.”20 The Red Desert indisputably contributed to Wyoming’s 332

reputation for superb horsemanship on the range and in the arena. There were several horse trap designs, a favorite being a box canyon partially fenced off at the front with wire or cable and sometimes disguised with brush. Other traps were brush or corral wings built out from a cliff or stone wall. Some were disguised corrals built around used water holes, and at night when the horses came in to drink, one or two concealed cowboys would leap up and close the exit, biding their time until the horses had drunk their fill. Sometimes the catchers used decoy horses to lure the wild ones into a trap.21 To get the horses from the trap to the railroad car or home corral, the horse catchers often used inhumane methods, some reluctantly described by John R. Barney. These procedures included closing the horses’ nostrils with “hog rings” so they could barely breathe and therefore not run; sewing up the animals’ nostrils with rawhide or baling wire; side hobbling them; tying them together at the neck while they faced opposite directions; attaching a heavy fifteen-foot chain (a “whip chain”) to one foot; tying a metal nut onto the forelock, causing the nut to bounce and make a painful sore; cutting a tendon at the hock; hog-tying several animals and coming back at leisure to collect them; or shooting the horse in the gut, which did not kill outright but let the animal live until it reached the chicken feed stop.22 In later years John Barney wrote, “Catching horses was a business for some of those fellows and a way of making a living. A lot of them felt they had to do whatever it took, to catch and sell horses, and they didn’t care as much about what they was doing to the horses at the time. They figured they was going to be killed anyway.”23 Horse catching was not just a summer occupation. William B. Logan Jr. said: “Sometimes they ran wild horses in the wintertime too. . . . [T]hey had grain fed horses to ride . . and the wild horses were having to forage for their food and fight the snow. . . . You could follow behind a trail that they broke for you, your horse didn’t have to work as hard. So quite often, in the wintertime, they did quite a bit of wild horse chasing.”24

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 333

Once caught, the horses had to be driven to a rail town for shipment. Frank Robbins, the famed horse catcher from Glenrock who worked the Red Desert for decades, was partial to Wamsutter, where there were pens in which the horses could be held for a time. Frank “Wild Horse” Robbins claimed to have taken 30,000 horses from the Red Desert in his career. Many went to rodeos in Rawlins or the annual wild horse rodeos held at Robbins’s Glenrock ranch. The Rawlins rodeo in the 1930s was big-time, the biggest between Calgary and Cheyenne. Robbins operated for decades in the northern half of the Red Desert, beginning in 1935. At one point he “made a deal with the grazing association to rid the range of the wild horses” by capture after he heard the association planned to shoot them.25 Robbins started his career near Flat Spring, where horses came to drink. He built corrals at Eagle Nest Draw, Sand Lake, Chalk Buttes, and the Chain Lakes. He pioneered the use of small aircraft in Wyoming to round up horses, believing this the most humane method of catching them, and worked with many regional pilots —Roy Lamoreau, Everett Hogan, Clyde Ice, Walt Williams, Howard Schrum, and Curly Wetzel.26 He was partial to the Chain Lakes area, known for its roan-buckskins. Robbins made several notable captures. One prize was a unique wild palomino he named Desert Dust. This horse was a publicity agent’s dream, featured in the 1946 film Fight of the Wild Stallions, shot in the Red Desert during Robbins’s annual horse roundup. The film shows in considerable detail the method of herding horses into canyon traps with the use of an airplane and featured some extraordinary camera work.27 Wild horses from the Red Desert went to fox and mink farms and to pet food producers. Some became ranch horses prized by the men who rode them. The less tractable bucked their way around the rodeo circuit as rough stock. Undoubtedly many went to horse dealers who, in turn, sold them to the U.S. Army for cavalry use. The final glory days of the U.S. Army Cavalry may have been Theodore Roosevelt’s charge on Texas horses up Cuba’s San Juan

Hill in 1898 during the Spanish-American War. During World War I both motor vehicles and airplanes were changing warfare. But the day of the horse was not over. More than eight million horses perished in the Great War, and although some were used in the front lines, most hauled ammunition and supplies to the front in horrendous conditions and over terrain where motorized vehicles could not go. The greatest number of horse casualties was from exhaustion and disease. There exist photographs of horses up to their knees in quagmires of mud, straining to pull huge guns and heavy loads of ammunition while under fire. In one photograph, three horses lie dead on the ground, all killed by the same shell, while their driver buries his face in his hands and sobs. Nor could exhausted horses rest: They were worn and weary and they should have been resting, but under the existing conditions real rest was impossible for most of them. It had been raining or snowing for days, and the ground was soaked with water. The teamsters of a fortunate few had found hard standing for their animals—a bit of sheltered road, the cellar of a ruined house, boards or tiles torn from a broken building perhaps—and it is surprising what hardships horses can live through if only they have something solid to stand on when off duty. But most of them were quartered on soggy earth which under their own trampling soon became a quagmire into which their legs sank deep. In vain a horse would draw up one hind leg from the sucking mud; the other would sink in just as far, to be pulled out in its turn a moment later.28 The United States exported more than 900,000 horses and 200,000 mules to Europe during the First World War.29 Some Wyoming ranchers, sensitive to the horses’ freedom, refused to sell to the Army dealers: “[T]he army approached Jim Hansen to buy cavalry horses. Hansen rounded up about 1000 head and brought them to the stock yards at Rawlins for delivery. When the army representatives came for the horses Jim asked, ‘What is to

Horse Bands of the Red Desert

333

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 334

be done with these horses? Will they go into the battle lines?’ The answer was ‘yes.’ Then Hansen opened the corral gates and turned all of the horses out to go back to the open range on the Red Desert. The army had to go elsewhere for horses.”30 As World War II began, some combatants still used horses. Poland’s tragic dependence on its cavalry of 86,000 horses was a terrible lesson that the days of the traditional warhorse were limited. In 1939, German dive bombers decimated the Pomeranian Cavalry Brigade, killing 2,000 of 3,000 horses. But the Germans themselves used 200,000 horses in the Polish campaign. According to some sources, the English and French, like the Americans, had let their cavalry languish in the period between wars while the Germans built up enormous herds of horses. It was not only mechanization and machine guns that defeated the Poles, but overwhelming numbers of horses, as well as stunning coordination of the German horse units with motorized vehicles and airpower. The Russians also clung to cavalry and horse artillery and depended on 800,000 draft horses to move heavy equipment. Yet in 1941 the Forty-fourth Mongolian Cavalry Division attacked a German infantry division, and within minutes two thousand Russian horses lay still or writhing on the ground. There were no German casualties. The Germans used horses not only in Poland, but in the 1940 offensive against Belgium, following up the cavalry assault with the lethal panzer tank divisions.31 Said Colonel Hardy, chief of the Army Remount Service, “In the campaigns in Flanders and France the German Armies included 791,000 horses.”32 In World War II the U.S. Army did not go to individual ranchers to buy horses as it had twenty years before. Wrote Colonel Hardy, “In a big buying program like this, we cannot possibly go from farm to farm and buy the number we want within the required time, so we are also buying from dealers now, and we always will have to do that in an emergency, but in normal times we buy from the man who raises them because we want him to get the maximum profit.”33 It is difficult to know how many horses from the Red Desert went to the European wars, but certainly 334

some ended their days far from the arid basin and sagebrush steppe where they were born. Wyoming people were intensely proud of their horses and their horsemen in World War I, but a generation later that had changed. In 1941 the 115th “Powder River” Cavalry Regiment of 1,295 men from all over the state made up the full complement of the Wyoming National Guard, then under federal command. Again and again the regiment was broken into smaller units, men transferred out, and the shrinking regiment ordered to one camp after another all around the country. At last only three troops still had their horses. In the summer of 1942 those horses were removed.34 The Wyoming men did not take it well: “In June, when the horses were taken away from the cavalry, the commander had to be confined to his office and the officers and soldiers had to be confined to their quarter to prevent a confrontation. But the horses were gone and so too was gone what the horses represented.”35 The Wyoming papers gave only passing attention to the swap of the horses for motorized vehicles. Something was indeed gone, and the Wyoming population was already sliding into a new set of middle-class values that touted automobiles and consumerism and relegated the horse to landscape decor. In the 1950s, Nevadan Velma Johnston, a.k.a. Wild Horse Annie, protesting the brutality of feral horse capture methods, became a passionate advocate for wild horses and burros. In 1959, Congress passed the Wild Horse Annie Act, prohibiting the use of planes and motorized vehicles to gather wild horses on federal land.36 In the next decade, two wild horse refuges were created, one on Nellis Air Force Base in Nevada, and the Pryor Mountain Wild Horse Range in the northern Big Horns. In 1971 the Wild Free-Roaming Horses and Burros Act gave the animals full protection on Bureau of Land Management and U.S. Forest Service lands— on paper. The statute came into conflict with ranchers who had livestock grazing permits on public lands. Some ranchers wanted to— did, and sometimes still do—shoot these horses that compete with their stock for grass. For others the horses were living symbols

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 335

of freedom, the natural world, and the historic past. The two views were powerfully opposed. The horse adoption program was not a great success. In the 1980s and early 1990s some of those “adopted” horses were sold for a profit to slaughterhouses. In 1997, legislation mandated that those who adopt a wild horse sign an affidavit stating they will not sell the animal for slaughter. There was also a requirement that adopters keep the horse(s) for one year. In 1957 a Wyoming rancher, Robert Brislawn in Crook County near Devils Tower, started a group dedicated to preservation of the mustang through careful breeding. He called his group the Spanish Mustang Registry and was soon nicknamed “Mr. Mustang.” Brislawn was deeply familiar with Wyoming mustangs. He rode many when he worked for the topographic branch of the U.S. Geological Survey in the years after 1911. When the mustangs he admired began to disappear, he started to gather up what he could find.37 The center of free-roaming horse society is a band, composed of a stallion and his mares, foals, colts, and yearlings. All free mares, unlike domesticated mares, come into heat between April and July, one month after foaling. Usually a trusted and wise mare leads the group, the stallion herding from behind, watching his charges and on the lookout for danger and mare-stealing rivals. Even today many stallions will duck and disappear, melting into the landscape, when a vehicle or mounted horses appear. Like the merlin, the Red Desert stallion has an almost magical ability to vanish from sight. Jack Price remarked that if the band was being chased hard, often the stallion would take up a front position, leading the escape rather than herding. There are, as well, bachelor bands of young male studs without harems, and occasional lone or duo horses. Young males are driven out of the band by the father stallion when they are three or four years old or younger. They run in teenage gangs, fighting and maturing until, at around ages six or seven, they are experienced and strong enough to build their own bands. The stallion must be constantly vigilant for interlopers that want to steal his

mares. Dominance and territorial protection rather than sex seem to be the driving impulses in the free stallion’s psyche. For management purposes the Bureau of Land Management (blm) has divided the Red Desert bands into different groups and herd management areas (hmas), tagged with geographical names such as White Mountain, Little Colorado Desert, Divide Basin, Salt Wells, and Adobe Town. Although most people regard the entire Red Desert area as blm land, perhaps because it is the largest remaining unfenced area in the contiguous United States and visually complete, it is really a complex patchwork of blm lands, Bureau of Reclamation lands, Forest Service holdings, Wyoming state lands, and privately owned land. Much of the privately owned land belongs to members of the Rock Springs Grazing Association, which, guided by John and Leonard Hay, descendants of pioneer Rock Springs settlers, did much to preserve the wild character of the region to the benefit of native fauna, including desert elk and pronghorn as well as small animals. Many Wyoming residents have strong feelings for wildlife and open spaces, epitomized by the character of the Red Desert region, where the old, hard identity of the state still lingers. Although the horse catchers, except for annual blm helicopter-and-horse roundups, are gone, the herds of the Red Desert continue to have problems. Generally strong and healthy, they are sometimes injured and sometimes ill, and there are too many of them.38 Horse numbers in general as well as horses too old for adoption have been an expensive headache for years. The federal government maintains an 18,000-acre retirement ranch outside Bartlesville, Oklahoma, for aging wild horses. Wild horses and burros multiply vigorously, at a rate that can double the population about every four years. There are roughly 20,000 wild horses and burros in sanctuaries and government holdings, and perhaps as many as 36,000 more running free on blm land. Pat Shea, the blm director under Clinton, remarked, “These horses are truly a great story of institutional resistance. No one has the gumption to actually deal with them.”39

Horse Bands of the Red Desert

335

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 336

In December of 2004, President George W. Bush signed an amendment to the 1971 Wild Free-Roaming Horses and Burros Act.40 The new law still lets wild horses and burros be adopted, but it also allows those over ten years old or that have been offered for adoption three times without any takers to be sold to whoever bids on them, including slaughterhouse representatives. The Bureau of Land Management believes this will reduce the number of free-roaming animals.

336

Notes 1. J. Frank Dobie, The Mustangs (1934; reprint, University of Texas Press, Austin, 1990), 23. 2. Edward Norris Wentworth, America’s Sheep Trails (Iowa State College Press, Ames, 1948), 128. 3. See Dudley Gardner, “Horses Come to the Red Desert,” this volume. 4. Settlers from the east also apparently believed that “civilized” eastern horses were superior to mustang stock. 5. Don Worcester, The Spanish Mustang (Texas Western Press/University of Texas, El Paso, 1986), 74. 6. Dobie, Mustangs, 321. 7. Tori Adams, “Spanish Horses’ Ancient Lineage Is Proven,” Rawlins Daily Times, February 25, 2006, 1, 4. Also see Patricia Mabee Fazio, “The Fight to Save a Memory: Creation of the Pryor Mountain Wild Horse Range,” Annals of Wyoming 69, no. 2 (Spring 1997): 28 – 47, for an overview of the Pryor Mountain horse bands of northern Wyoming– southern Montana and the humans interested in them. 8. Jack Price, Wild Horse Country in Wyoming (Pine Hill Press, n.p., 1996), 148 –149. 9. John F. Gooldy, “Early Day History of the Little Snake River Valley” (unpublished document, American Heritage Center, ca. 1962), 2. Cowboy John Barney described the plight of a band of range horses in a hard winter of the late 1920s: “That winter we got a lot of cold, stormy weather and the snow got awful deep. Around that country [the southeast edge of the Red Desert] there was a lot of open-range, so horses by the hundreds drifted in and all around the ranches and lived on the hay that was spilled by the ranchers. . . . [T]here would always be a big bunch of horses that would come in looking for something to eat. My boss . . . wanted me to shut up all the gates on the ranch and gather up all the range-horses. . . . It was snowing so hard that I couldn’t hardly see my hand in front of me. I rounded them up and put them through the gate. I got off my horse and shut the gate. [T]he wild horses just stood there and didn’t want to leave. Their manes and tails were real long and their foretops were hanging over their eyes. I saw that some of them were already getting thin and they sure looked hungry. . . . I leaned over in the saddle and just sobbed and sobbed for a few minutes because I felt so sorry for them horses. I knew as hard as the winter was that most of them wouldn’t be around much longer. As I rode away they just stood and watched me like they hoped I might be able to feed them. I’ll tell

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 337

you, it was real hard for me to leave them horses like that.” John Rodney Barney, “Looking Back”: My True Memories from 1908 –1996 as told by a “Real Cowboy” (Commerce Printing, Sacramento, Calif., 1999), 58 –59. 10. William B. Logan Jr. interview by Shirley Black, n.d., Western Wyoming Community College Oral History Project (hereafter “WWCC Oral History Project”), 1:7. By “Hamiltonian,” Logan likely meant “Hambletonian.” 11. Helen M. Morgan, Snake River Profiles (Baggs, Wyo., 1970 –1972), 2:20. 12. Ibid., 3:7– 8. Helen Morgan lists Pete Olsen as a horse runner and bronco rider who ranched north of Rawlins. He in turn named Frank and Tode Espie, Al Props (Desert Rat), Boots Young, Fat Heady, Jim Karstoff, Ruby Ortiz, Albert Robinson, Buck Buchanan, Phil Espie, and Fred Rasmussen as “fellow horse-catchers.” In the mid-1920s, Morgan adds that Ernest Long and Guy McNurlan ran wild horses around Lookout Mountain. Others were “Shorty Creol, Babe McCarger, Bert Cole, Buck Ayer and Punch Ellis.” 13. William B. Logan Jr. interview, 7– 8. 14. L. H. Chivington, “Last Guard”: Memories of an Old Time Cowboy from 1901 to 1926 in Northwestern Colorado (unpublished manuscript, Museum of Northwest Colorado, Craig, n.d.), 183–184. Although Chivington identified with northwest Colorado, he cowboyed in Wyoming’s Red Desert and in the Little Snake River valley. This is one of the few accounts of the work of a “common cowhand” in and around Wyoming from 1900 to the early 1920s. 15. Gordon G. Kinkaid, interviewed by Rhonda Kinkaid in 1988 for the WWCC Oral History Project, 1:2, remarked that his parents bought a ranch “from either Latham or Teeters. . . . It was a horse ranch. . . . About 1908-10-12 along in there, . . . your Model T’s came into existence so the horse business was ridiculous. . . . [T]hey kept them [the horses] but you could never sell any of them, and they weren’t worth nothing, nobody wanted horses. . . . But he kept them until 1934 and then they shipped everything they could gather . . . they rounded up all they could find and shipped them and sold them for dog food at that time.” 16. From Walker D. Wyman, The Wild Horse of the West (University of Nebraska Press, Lincoln, 1945), 204: “By designating a carload of horses as ‘chicken feed,’ the railroad was under no legal obligations to give humane treatment to the cargo. Under this rate thousands of horses purchased for one cent a pound or less on ranges as far east as the Dakotas, were transported to California. This was almost the only market for scrub horses caught by cooperating cattlemen, mustangers, and one or two

government bureaus in the 1920’s.” 17. Price, Wild Horse Country, 172, remarks that his father sold horses to the Black Willow Mink Ranch in Coalville, Utah. Helen Morgan, Snake River Profiles, 3:7, says, “The Chapel brothers from Rockford, Illinois, came and bought the horses for dog and fox food. They gave ten dollars a head for the horses over 1,000 pounds and five for under, guessing the weight.” 18. John Rolfe Burroughs, Where the Old West Stayed Young (Bonanza Books, 1962), 173–174. 19. John Rodney Barney, “Looking Back”: My True Memories from 1908 – 1996 (Sacramento, Calif., 1996, 1999), 137. Some white wild birds, such as pelicans, have black primary feathers, as melanin stands harder wear. On the hardness of hooves, Jack Price, Wild Horse Robbins (Pine Hill Press, Freeman, S.Dak., 1999), 15–16, also says, “[T]he type of terrain and soil have a lot to do with the formation of the feet. Colts born on rocky, hard ground develop small, hard feet, as opposed to those born and raised in sandy, soft soil. Horses growing up in the sand dunes have a tendency to have flat feet, giving them a definite advantage for running in the sand. The land around Dobe Town . . . is very rocky and rough, and helped to develop a nearly perfect horse.” 20. William B. Logan Jr. interview, 7. 21. Price, Wild Horse Country, 130, describes the construction of a winged corral. 22. Price gives details of sewing up the nostrils: “They got their nostrils sewn shut with one strand of buckskin. They weren’t tight enough to shut off breathing, but restricted it when they ran. . . . This seemed like quite an inhumane treatment until I saw it done. They just made a small incision in the outer edge of each nostril and put the thick buckskin string through and drew it part way shut” (ibid., 144). 23. Barney, “Looking Back,” 136. 24. William B. Logan Jr. interview, 7. 25. Ibid., 52. One assumes this was the Rock Springs Grazing Association. Robbins reputedly made an impassioned plea for capturing rather than shooting horses. 26. Ibid., 71. 27. In 2004 Frank Robbins and Desert Dust entered the pantheon of Wyoming cowboy-and-horse heroes. Both are the subjects of a series of murals by Missouri artist Kerry Hansen, planned for the Ferguson Building on Fourth Street in Rawlins. 28. Ernest Harold Baynes, Animal Heroes of the Great War (Macmillan, New York, 1925), 28.

Horse Bands of the Red Desert

337

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 338

29. Col. E. N. Hardy, Horses and Mules in Modern Warfare (Horse and Mule Association, 1941), 8. 30. Louise Bruning Erb, Ann Bruning Brown, and Gilberta Bruning Hughes, The Bridger Pass Overland Trail 1862 –1869: Through Colorado and Wyoming and Cross Roads at the Rawlins-Baggs Stage Road in Wyoming (ERBGEM, Littleton, Colo., 1989), 114. 31. Elwyn Hartley Edwards, The New Encyclopedia of the Horse (DK, London, 2000), 324–325. 32. Hardy, Horses and Mules, 10. 33. Ibid., 15. 34. T. A. Larson, Wyoming’s War Years 1941–1945 (Stanford University Press, Stanford, Calif., 1954), 15–16. 35. Michael Cassity, “‘In a Narrow Grave’: World War II and the Subjugation of Wyoming,” Annals of Wyoming 68, no. 2 (Spring 1996), 4. This article is a valuable counterweight to Larson’s more progressoriented review of the changes World War II forced on Wyoming. 36. Patricia Fazio, “Fight to Save a Memory: Creation of the Pryor Mountain Wild Horse Range (1968) and Evolving Federal Wild Horse Protection through 1971” (Ph.D. diss., Texas A&M University, in Annals of Wyoming 69, no. 2 [1997]). 37. Worcester, Spanish Mustang, 76 –77. 38. In the autumn of 2004 Charles Ferguson, the geologist working on this Red Desert project, was camping on Powder Rim above Shell Creek. He recalled: “The last day, I did some mapping there and noticed a lone white horse down along the creek that did not move for the longest time. I finally checked it out with my binoculars. It seemed to be resting or asleep, but it finally moved slightly. It had a bunch of magpies flittering about and walking all over its back. I kept watching and then it lifted its right front leg as if to paw the ground or step forward, and I could see as the leg lifted up above the sage brush that the lower leg was a bloody stump. There were more magpies on the ground. Coyotes had been yipping down there the night before. Must have chased it and it may have snapped its leg in a hole. Perhaps the coyotes attacked and pulled the flopping lower leg off. Very gruesome” (pers. comm.). 39. Martha Mendoza and Angie Wagner, “Too Many Horses on the Range,” Associated Press, in Rawlins Daily Times, July 21, 2004. 40. Adopters had to sign an affidavit stating that they would not sell their adopted animals for slaughter.

338

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 339

25. Opening the Oyster Annie Proulx

ne way to look at the history of Wyoming is to sort the past into periods of dominant activity: exploration and fur trapping; emigrant trail passages; the decades of Indian wars; railroad and telegraph building; the rise of the cattle and sheep industries; mining and territorial governance. The last hundred-plus years since statehood have been the time of extraction of natural resources and raw materials, with efforts to diversify the economy through tourism and encouragement of small businesses. In all of these stages the Red Desert has been important to humans, and today, in the current energy boom, it is a central extraction site. Jefferson’s Louisiana Purchase of 1803 brought vast regions under the flag, but the derisive term “Great American Desert,” which referred not only to specific deserts but the entire west, was the popular view and few people thought any part of the west worth a second glance until political circumstance and the entrepreneurs Manuel Lisa, John Jacob Astor, and William Henry Ashley illustrated the possibilities for wealth in the fur trade.

O

Ashley William Henry Ashley, a risk-taking frontier businessman and Missouri political figure, with his partner, Andrew Henry, revolutionized the western fur trapping business through good or-

ganization and bypassing the traditional trading post and Indian fur-gatherer relationship. Ashley introduced a commission system, set up a quasi-military management structure that still allowed trappers independence, and used advertising, cultivated the media, and stroked politicians to gain advantages.1 His famous want ad in the Missouri papers for “one hundred young men to ascend the Missouri River to its source, there to be employed for one, two, or three years”2 was answered by FrenchCanadian boatmen and by others who later became great names in the Rocky Mountain fur business—Hugh Glass, Jim Bridger, Jedediah Smith, William and Milton Sublette, James Clyman, Tom Fitzpatrick, David Jackson. Ashley was self-disciplined and inscrutable, with an unusual flair for organization.3 He was involved in mining lead, manufacturing gunpowder, and speculating in land. Shipwreck, Indian war, duels, lawsuits, economic depression, debts, and fluctuating markets jerked his business affairs around like a rat in a terrier’s jaws. But he had a nose for opportunities. When fur prices rose around 1821, Ashley, whose partner in the lead mine, Andrew Henry, was an experienced trapper, sensed that important money could be made in the west. He set to work getting the financing together. Ashley’s credit and reputation were good, and he raised the money to support a fur trapping expedition of 125 men for 339

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 340

Geophysical Operations in the Red Desert russel l. tanner

From almost any promontory in the Red Desert, miles and miles of straight, seemingly endless paths are visible across the landscape. These are the result of geophysical exploration operations in the 1950 and 1960s. In those days it was common practice to blade a line, usually along township or range lines, with a bulldozer or road grader, clearing a path for drilling rigs to conduct geophysical exploration. At that time there was no regulation of such operations on federal lands. The bladed path facilitated travel of large pieces of equipment, including a small drilling rig, fuel trucks, and trucks hauling heavy receiver cable along the transect being studied. The operators drilled rather shallow — roughly 100-foot-deep —holes loaded with explosives wired together so they could be set off simultaneously or at specific intervals as in mine

blasting operations. Heavy cable with receiver devices attached set at specific intervals recorded and timed the upward reflection of the explosive blast as it echoed from the deep stratigraphic formations. The reflected shock took longer to return to the surface through dense strata than through less dense layers of rock. This method of using explosives in drilled holes is commonly called shot-hole geophysical, since the explosive charge is packed in a hole in much the same way that black powder is loaded into the barrel of a flintlock or percussion cap rifle or shotgun. These bladed paths have left virtually permanent scars across the landscape of places like the Red Desert. From the air these paths can be seen almost everywhere in western North America and likely many other places in the world. Those early studies provided baseline information that aided in the discovery of oil and natural gas-bearing geological formations. One

several years. The first boat, loaded with supplies, sank three hundred miles upstream, and another had to be outfitted at considerable expense. In 1822, the lost boat and its goods replaced, Ashley and Henry and their green trappers headed up the Mississippi, intending to travel by water to the mouth of the Yellowstone, then switch to horses and make their way west. They ran afoul of the Arikaras on the upper Mississippi, and the battle that followed, in which Ashley lost men, horses, and supplies, led to the first U.S. government attack on Indians beyond the Mississippi, a venture fumbled by the commandant, Colonel Henry Leavenworth.4 Ashley and Henry reviewed their losses and, in the face of great odds, decided to press on overland. Ashley went back to St. Louis to raise more money for 340

limiting factor in the applicability of data gathered from these operations is that they provided only a two-dimensional view of the deep geological strata. Two-dimensional operations continue to be used, though generally over much smaller areas, to refine the geologist’s knowledge of deep geological structure. Since about 1990 this two-dimensional method of studying subsurface geological stratigraphy has been replaced by techniques that give geologists a three-dimensional picture of deep strata. Three-dimensional geophysical provides a view of buried stratigraphic levels and measures their potential mineral deposits— especially fluid mineral deposits (oil and natural gas). This method requires denser spacing of points (either shot holes with explosives or vibroseis placements) than the older two-dimensional techniques that relied on induction points roughly every 660 feet. Vibroseis uses a “thumper,” a metal pad driven by a diesel engine, to strike the ground, inducing a

the fur venture, and Henry continued on west. It was a hard trip for Andrew Henry. Wilderness life was a young man’s game. When he returned to Missouri the next year, he decided to retire. Ashley was now trembling on the brink of bankruptcy, his creditors twitching with anxiety. The fur venture seemed yet another loss in Ashley’s long experience of business reverses when a letter from Thomas Fitzpatrick arrived.5 In it Fitzpatrick said that he and Jedediah Smith’s party had enjoyed good trapping. They had followed the Crow trail through the South Pass on the northern edge of the Red Desert and entered the Green River valley, which teemed with beaver.6 Ashley immediately understood that his men had cached enough pelts to put him out of debt and that there was the gleam of the chance that

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 341

shock wave that is measured to reflect deep geological structures. Three-dimensional geophysical operations, according to oil company geologists, is much more effective in mapping the geological structures likely to hold hydrocarbon reserves in the form of natural gas or oil than the old explosive shot-hole method. The first few vibroseis operations in southwestern Wyoming used three thumper trucks in a straight line to drive across the landscape and strike vibration points about every 330 feet. Bureau of Land Management (BLM) officials quickly noted that this approach left a scar on the land nearly as pronounced as the old bladed paths. Thereafter, the agency required the thumper trucks to be spread out so there were three adjacent paths rather than a single route used by the three trucks. Geophysical operators also brought in trucks equipped with large balloontype tires to lessen the impact of the vehicles’ weight by dispersing it across a larger surface

area. This same type of tire is used in natural gas development operations on permafrost above the Arctic Circle. While it has been argued by some environmentalists that vibroseis is a more detrimental exploration technique than the old shot-hole explosive methods, BLM specialists who have field-checked vibroseis operations areas both from ground level and from low-level helicopter overflights have determined that the new strategy of employing balloon tires on vibroseis trucks spread out, rather than in line, creates less surface disturbance than any of the older shot-hole geophysical technology. Furthermore, the information gained from these studies can much more accurately pinpoint gas and oil reserves, thus resulting in drilling far fewer dry holes, or unproductive wells. Three-dimensional geophysical operations have been conducted over thousands and thousands of square miles of southwest Wyoming, and especially in the Red Desert,

a good winter trapping season would make his fortune. In 1825, without a partner who could go into the field, Ashley himself accompanied the year’s expedition. His plan was to pick out a site for an annual rendezvous, which would eliminate the traditional trading post and, at no construction cost, allow trade and barter between free trappers, various tribes, fur buyers, and traders. Whether the idea came from his own observation of annual Indian trade fairs or from the suggestion of one of his trappers is not known. He set the first rendezvous at the mouth of Henry’s Fork in what is now southwestern Wyoming; then, when the gathering time arrived, surprised by mosquitoes, he shifted twenty miles upstream to a location between Burnt Fork and Birch Creek near Brown’s Hole.7 In subsequent years the ren-

with minimal long-term damage to the landscape. Other technological improvements, including application to global positioning system (GPS) surveying methods, and use of data chips to lessen the amount of cable needed to record vibration reflections, have also lessened the footprint of geophysical operations on the ground. Nonetheless, geophysical operations are just the tip of the iceberg of the effects of gas and oil operations in high desert regions. Although the scars still visible from fifty-year-old geophysical operations remain, and newer exploration strategies are less harmful, the net effect of full field development, especially of natural gas deposits, over vast regions is stamping the landscape with permanent industrialization scars of access roads, drilling locations, and pipelines.

dezvous locations shifted from one riparian area to another. Ashley floated the upper section of the Green River in a crude boat of hasty streamside manufacture before John Wesley Powell (1834–1902), much celebrated for his descent of the Colorado in 1868, was born. The mid-1840s, when the fur trapping business collapsed, were years of astonishing territorial gain for the United States, all of it new territory west of the Mississippi, terra incognita to all but the trappers and traders who had been working the country for decades. Now every white person entering the country could regard himself as an explorer.8 The U.S. Regular Army rode high after its victory in the Mexican War. In 1845, Texas was annexed. The long-simmering Oregon boundary dispute between Britain

Opening the Oyster

341

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 342

and the United States, which often took the form of fierce rivalry between the men who worked for the Hudson’s Bay Company and North West Company and the American free trappers, ended in 1846 as the fur trade waned. The peace settlement of 1848 (Guadalupe Hidalgo) with Mexico ceded California, the Rio Grande country, and the Great Basin region to the United States. Suddenly the United States territory spanned the continent. It was well known that west of the Missouri and Arkansas rivers lived thousands of Indians in many tribes and speaking many languages. With the newly acquired territory came Thomas Jefferson’s idea of “systematic Indian removal,” shifting tribes from the east to the west, using the new Louisiana Purchase territory as the repository for the relocated tribes.9 This Indian country was to be a “Permanent Indian Frontier,” marked by a line of military forts. Something was known of the new country and the Native inhabitants. For decades, since 1793 when Pedro Vial made clandestine trading trips from San Antonio and St. Louis to Santa Fe, and especially since 1821 when the Santa Fe Trail became a legitimate passageway, business entrepreneurs had carried fabrics and other manufactured goods to Santa Fe and Taos, in exchange for silver and mules in profitable trade. For decades these merchants had pressed the American government for expansion in the region. One of the dominant strands in the American character was and is the itch to make money, and the modus operandi in this period was mercantile trade and exploitation of natural resources, supported by the supply and transportation businesses. Thomas Jefferson’s instructions to Lewis and Clark expressed it clearly: “The object of your mission is to explore the Missouri river, & such principle stream of it, as by its course & communication with the waters of the Pacific Ocean, may offer the most direct & practicable water communication across the continent, for the purpose of commerce.”10 Southwest trappers used Santa Fe and Taos as bases. They took furs in the Southern Rockies, though historians traditionally have seen the beaver-rich tributary streams of the Green River 342

region as more important fur country during the 1820s and ’30s. The fur trappers, later romanticized as fearless, freedom-loving individualists, the very essence of American character, have more recently been characterized as Jacksonian men, entrepreneurial opportunists who saw a chance to make their fortunes, following the example of William Ashley, whose brilliant success showed the monetary possibilities in the fur trade.11 A few—William Sublette, Pierre Chouteau, even Jedediah Smith, who would not live long enough to retire— did make money, and their success was an encouragement to hard-muscled country boys whose ambitions reached beyond their fathers’ hardscrabble farms. If a man had guts and quick responses, there was a chance to make a small fortune in a few seasons, and if he had brains as well, he might climb up the fur business hierarchy as did Ramsay Crooks.12 But there was another trapper type—the withdrawn “westering man,” exemplified by Joe Walker. Walker and others like him melted like smoke into the wild country and the roving life. The rich years of the beaver trade, the specialization of the men who trapped the Green River region, were brief, from the 1820s and rapidly fading by 1840, when nutria fur and silk replaced beaver as the raw material for British and European men’s hats.13 The trapper’s life was wonderful for certain men—the chance of making a pile of money, plenty of sex in a period of Victorian repression, the adrenalin rush of combat with a wily foe, the mental and psychological pleasure of exploration, mapping and learning new country, the fellowship of like-minded men, the exhilaration of competition with rival trapper outfits and companies, and the mad debauches of the later rendezvous. Perhaps it was the ideal masculine life, and no wonder that envious male historians have romanticized the mountain men and that the trappers became major figures in the American myth of adventure and wilderness conquest. Although many of the trappers had long and apparently happy liaisons with Indian women — the so-called country marriages that gave them a business advantage through the woman’s family

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 343

and tribal connections—these men also set the behavioral pattern of exploitation and hostility that characterized later relations between the United States and the western tribes, forcing what David Lavender called “the tragic choices which the Indians repeatedly had to make in struggles they could not possibly win.”14 Emigrant Trails Energetic and competitive, Americans have always reacted to a chance for the quick fortune, and when the news that gold had showed in the gravels of the American River in California reached the east in August of 1848, it set off a terrific and unprecedented stampede of people pushing their way to California by any and every available means—by ship, by the Butterfield and CaliforniaOregon-Mormon trails, by the Cherokee and Overland trails through the Red Desert region. That rush to gold was like the rupturing of an earthen dam that shows first a small wet spot like a shadow on its face, then a pencil-size spout, then a fire hose stream, and all at once an unstoppable torrent of water, rock, mud, sand, fish. Here began the flood of human beings out of the strictured east and into the open west, bursting through the imaginary line of the Permanent Indian Frontier and into tribal hunting grounds. The whites followed ancient trails, especially the CaliforniaOregon-Mormon Trail, which many still believe the fur trappers made but which was essentially a retracing of Indian trails that themselves followed game trails, particularly those of bison, formed over thousands of years by migrating beasts. Early historians of the west made a fuss over the “discovery” of South Pass and the Cheyenne Pass, and their strained efforts to affix dates and the names of white “pathfinders” to these gateways, well known to tribes who often acted as guides to the invaders, flavors their research with nationalism, justification, and the guilt of the invader. Coal, Oil, Trona, Uranium, and Methane Gas Occasionally, traveling through and around the Red Desert, one

glimpses a manufacturing or processing plant, usually in a location where few people notice it. Some of these are refineries, process trona or uranium, or extract sulfur from natural gas. Coal, oil, and gas, naturally occurring substances in the earth, the nonrenewable fossil fuels, are not minerals but transformed organic material. All of them are largely made up of two elements, hydrogen (a gas) and carbon (a solid). Uranium is a mineral found mixed in soils and rocks. These subterranean substances, so plentiful under Wyoming sageland and bunchgrass cover, and so attractive to outside energy corporations, have led the state in a vigorous dance of boom-and-bust over the past century and earned it the sobriquet of economic colony. But in the 1960s the state enacted an oil and mineral severance tax and set up the Permanent Wyoming Mineral Trust Fund. The severance tax was the culmination of a proposal that had been kicking around since the 1920s. In recent years the tax has made the state wealthy, with more than $3 billion set aside in trust funds, and cautious apportionment of the income is reshaping the state’s character. On the other side of the rosy picture, energy extraction is having a negative effect on what many once considered the state’s most precious assets— clear air, world-class scenery, and extraordinarily rich and diverse wildlife. coal Coal comes from vegetation buried during the Carboniferous period 360 –290 million years in the past. As organic material— leaves, trees, grass, soil— decomposed over the millennia, a stupendous compost pile compressed by more and more deposits, it first became peat (60 percent carbon), then “brown coal” or lignite, then soft bituminous coal (80 percent carbon), and, finally, hard, shiny anthracite that is 95 percent carbon. Wyoming has vast deposits of high-quality, low-sulfur coal thanks to its ancient swamps. Those coal deposits made the nineteenth-century Union Pacific’s development and operation possible. The state was literally founded on coal, and coal continues to be of great importance

Opening the Oyster

343

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 344

despite the belief of many scientists involved in climate change studies that the release of anthropogenic carbon dioxide into the atmosphere is pushing the earth out of atmospheric balance. Most coal-burning power plants convert water to steam, which turns the huge turbines that create electricity. This antique technology, inefficient and polluting, is still used, although newer technologies, none perfect—such as integrated gasification combined cycle (igcc) power plants that use a catalyst to draw gases from the coal without burning it—are being built in some parts of the country. Early coal extraction in Wyoming was largely underground mines, the coal firing Union Pacific’s locomotives. When diesel engines arrived in the 1950s, many coal towns folded up, and underground mines gave way to surface mining of coal beds, most of the coal shipped out of the state by coal trains to electric power plants all over the country. This was the beginning of the so-called energy era, and Wyoming, rich in natural energy resources, though it clung to its old identity as the cowboy state, was, in truth, the energy state. There are power plants in the Red Desert. The Jim Bridger Power Plant opened in 1974 east of Point of Rocks, adjacent to a vast supply of coal. The Jim Bridger is a 2,110-megawatt generating plant that supplies six western states with power by way of PacifiCorp and Idaho Power. Each year, the plant uses nine million tons of coal from the nearby Bridger Coal Mine. That source is dwindling, and in 2004 Bridger applied for permission to mine the Ten Mile Rim (on public BLM land), proposing the use of the longwall technique, which sends remote-controlled continuous miner machines down to tunnel through the seam. Although the coal mine towns of Rock Springs, Superior, and Hanna have received some attention, early Carbon, on the northeast fringe of the Red Desert and with a large Finnish population, was one of the first old-style mining towns, somewhat neglected by historians.15

344

oil Both oil and gas originate in the organic content of primordial marine or lacustrine sediments. The pressures and heat of increasing sedimentation force the organic molecules into complex chemical changes that become oil and gas. Oil and gas hydrocarbons grade out commercially into a gradual succession of gases, liquids, semisolids, and solids. On the light end of this scale is methane gas, odorless and flammable, then gasoline, naphtha, kerosene, diesel, and fuel oil, lubricating oils, wax oils, paraffin, and asphalt. The gases fall into two categories: dry, or fixed, gases, and wet gases. Wet gases, when compressed or chilled, make a volatile, very light gasoline referred to as natural gasoline, or drip gas. Oil is far less abundant than coal and harder to find. As all the world knows, the supply is limited. Oil from natural seeps has been used for medicine and illumination since prehistoric times. In Wyoming, Indians used seep oil as medicine. They shared their knowledge of the location of these seeps and pools with trappers and mountain men, and in 1832 at the rendezvous on the Green River, trappers mentioned the existence of an oil spring near South Pass to Captain Benjamin Louis Eulalie de Bonneville, trapper, trader, and adventurer. Travelers across Wyoming, including Stansbury with the Topographical Corps, early settlers, and army troops, all noticed oil seeps and springs. By the 1870s interested people eager for the curative balm were digging pits near Evanston. In 1883 the first flowing well in the state came in not far from Bonneville’s tar spring, part of what later became the still-producing Dallas field. Around the same time, several men became interested in the Salt Creek area north of Casper. By 1889 an editorial in the Casper Mail stated, “There are springs in various localities that flow all the way from one gallon to ten barrels a day. It is by the springs that the oil belt of Wyoming is traced for more than two hundred miles.”16

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 345

Old Carbon nancy anderson

On June 29, 1868, Union Pacific Railroad workers reached Carbon City—656 miles from Omaha, elevation 6,750 feet, and with valuable coal deposits.1 The camp lay at the eastern edge of the formidable Simpson Ridge. Thomas Wardell of Missouri contracted with the railroad to develop the Carbon mines. They discovered coal at eighty-five feet, built hoist and tipple. By September, Mine No. 1 was producing one hundred tons of coal a day.2 Wardell’s labor problems emerged even as the camp was established. In 1871, troops were called from nearby Fort Fred Steele to quell miners’ unrest over wages and the right to strike.3 In early 1874 a miners’ insurrection led to a lockout. A letter writer who signed himself “A. Miner” voiced his grievance: “It has been the habit of the company for years past, to employ about double the number of men needed, thus giving the miners about half work. This was done in order that to procure the necessaries of life, we were obliged to draw nearly our whole pay from the company store at which prices range from 50 to 100 percent higher than any other store in the country, and those who refused to trade at said store were discharged. . . . This is the reason for our present suspension.”4 Jay Gould dominated the Union Pacific in 1873 after buying up one-third of the stock. Within a month of A. Miner’s letter to the Independent, Gould snapped the contract with Wardell and took control of the mines, establishing the Union Pacific Coal Department, with D. O. Clark, “that stern human machine,”

at its head. Controlling both transportation and mining operations gave the Union Pacific a virtual monopoly.5 The predatory company store was one issue with Carbon’s miners; cheaper foreign labor and the definition of a ton were others. In Carbon the foreign-born generally were Finlanders. The Finns began arriving in the 1880s, and by 1900 Carbon had 140 Finlanders among its total population of 634 persons. Mont Hawthorne, a Carbon miner from 1881 to 1883, observed: “Most of the Finns had worked so hard in the old country that they was willing to work for less and put in as high as eighteen hours a day. . . . The Lankies, who was old hands at mining, got mad and said they was running down the pay for all of us. . . . The worst of it was that it took 2,240 pounds to make a ton. . . . The company took one car in eight for screenage and didn’t give you nothing for that car. Before long they changed that to one in four. The Union Pacific was selling that fine coal to the farmers in Nebraska . . . getting ten dollars a ton for it.”6 Seven mines operated during the lifetime of the camp, four of them within the narrow valley floor occupied by Carbon City. Carbon mines were relatively free of explosive gas, but there were other hazards—falls of rock, caveins, crushings by car, and falls down shafts.7 The foreign-born were prey to certain elements. Wyoming became a state on July 10, 1890; on August 13, seventy-two men, most of them Finlanders, received citizenship at the Carbon County Courthouse. On August 16, “Editor Friend” in the Democratic Carbon County Journal railed with some exaggeration: “Two hundred and sixty [sic] foreigners were

Opening the Oyster

rounded up at Carbon, Hanna, and Dana by Sheriff Hadsell and . . . brought for naturalization. . . . They are supposed to be republican, as the party in charge used every precaution to prevent those of known democratic tendencies from riding on the special train.”8 As Carbon changed from man camp to a village of families, it became “a busy little burg with much social activity.”9There were strawberry festivals. Men cleared a patch of sagebrush, and “[h]orses were raced there every Sunday, with young John and Michael Quealy, sons of the first Carbon pit boss, perched in the big-horned western saddles as jockeys.”10 Another favorite pastime was shooting live pigeons. A purse of $250 went to wrestler Jim Finch (trainer, Bobby Fogg) for defeating Finn Jacob Myllimaki (trainer, Matt Wakanomialersteinkoski). Baseball appeared in Carbon as early as July of 1878, when “a picked nine of the miners played nine of the day hands two innings.”11 Then, on the night of June 30, 1890, a fire consumed most of the town north of the tracks. A new, planned Carbon rose on the south side of the tracks. Town ordinances demanded brick chimneys and forbade naked stovepipes protruding from roofs. But the Union Pacific moved the main line north to Chimney Springs, renamed Hanna, and Carbon found itself on a spur. In 1902 the last working mine in Carbon closed. Mont Hawthorne was the poet of Carbon. Following a peripatetic father, the family had gone from a failing plantation in the south, to surviving starvation, grasshoppers, typhoid, and cloudbursts on a Nebraska homestead, to grading railbed in Colorado. In 1881 Mont, at

345

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 346

fifteen, was sent ahead to Carbon. He remarked later: “Nobody done no planning at Carbon. The town just growed out of the desert. Except for the regular buildings at the mine, and a few stores strung along down by the depot, everyman had to provide for hisself, and there wasn’t much to provide for with. There wasn’t a tree in sight, nothing growed at Carbon but sagebrush, and greasewood, and prickly pears. . . . Nothing was wasted at Carbon. Men fought over the packing cases around the stores and mines. . . . Blasting powder came packed in big tin cans . . . miners would take them home and flatten them out to make roofs for their homes.”12 The family arrived, and Mrs. Hawthorne set up housekeeping in a dugout, supplementing the mining wages with bread baking and nursing, a vital skill in the isolated camp where life was precarious. One night the miners come for Mama and she had to walk four mile through the deep snow to Number 5 mine. A fellow lay back on one of the lower levels where a coal car had run over his leg, breaking it up close to the hip, and cutting him bad. . . . She kept her fingers on his leg, feeling them bones slip back in place, while a couple of miners held him down by the shoulders and two others pulled down on the leg, to set the broken bones. . . . She seen he’d lost too much blood . . . and the miners brung what she needed and they set up a place right there in the corner of a shed for Mama to stay and take care of him. Nights, Myra would stand by the window watching more than a hundred miners come marching

346

over the hill . . . nothing showed but black shadows and bright lights, moving across, above the snow. And Myra would cry and wonder why Mama wasn’t coming home from Number 5 mine, too. Then came the night when we seen the shadow of her long full skirt as she walked along beside the men who was carrying somebody strapped to a plank.13 Mont Hawthorne credited storekeeper Coffee Johnson with inspiring him to move on.14 A Swede, named Coffee Johnson, had a regular museum in his store. Coffee was a great traveler. . . . Nights I’d go into his store and stand and look at the bottle of water from the Dead Sea, and sand from the Sahara and the bullets from Gettysburg, the lion skin from Africa and the eightlegged lamb that growed right here in Wyoming. That’s when I knowed it was time for me to be getting out where I could see new things for myself. . . . Living in Carbon in them days was like living on an island. The trains stopped for coal, then they went on again. Us folks mining that coal, didn’t have no way to turn, excepting towards each other. We come there to live on a desert. If we hadn’t put down roots and worked together, we couldn’t of had no more chance than a bunch of tumble weeds, blowed along ahead of the wind.15 Oskari Tokoi left Finland as a boy of seventeen, escaping “the Arctic wastes, the forbidding forests, the icy rivers, the desolation . . . the Great Hunger years of all killing frost.”16 For

Annie Proulx

the able-bodied men of his village, it was usual to travel to the outside, to the south, to the sea, even to America. Men who had never seen a mine in their own country became common miners who fueled the early Union Pacific Railway. Oskari arrived in Carbon and moved into the Pelanders’ Finnish boardinghouse, which, including others, he later described as “all in a row and half-dug into the hillside so that only their street fronts had windows. Each of them operated on the same principle as Pelanders.17 Carbon’s boardinghouses were probably the first Finnish cooperative undertakings in America, although histories of the cooperative movement make no mention of them.”18 Oskari joined both the American Good Templars League and the congregation of the Finnish Lutheran Church, where monthly services were held by an itinerant preacher in a small church built by the miners.19 Another sanctuary for the Carbon Finlanders was Arbor Valley, Lehtilaakso, a spring-fed slope: “This vale had become a Finnish shrine where they went to devoutly, as to a church . . . from which they would bring back a branch of aspen to set upon a shelf or a wall and to cherish until it became sere.”20 Within a year of his arrival, Carbon miners were cut to only two or three workdays a week. Search for better pay and conditions took him to other Wyoming coal mines —Almy, Inez, Glen Rock. As the depression of 1893 deepened, many young Finnish immigrants took to the road. During brief employment at the Homestake Gold Mine in Lead, South Dakota, Oskari noted, “I first came into contact with the American Labor movement which opened

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 347

my eyes to new forces in a new world and laid the foundation for much of my future activity.”21 Returning to Carbon, Oskari found that miners there were deserting the old camp. At nearby Hanna two hundred Finns were working, and more arriving each day. Hanna’s No. 1 mine was famous for the thickness of its coal seam and for its dangerous conditions. Two weeks after Oskari’s employment there, the miners were ordered out; there was a fire on the thirteenth level. Oskari described the chaotic exodus: “There was fearful crowding at the narrow exit. The men and mules from the thirteenth level were there first and were pushed from the rear by the men of the twelfth who were in turn shoved by those coming up from the still lower levels. The ‘manway’ exit was not only narrow but steep, making for slow and careful maneuvering. The confusion grew worse as the smoke thickened. Men were choking, praying and swearing.”22 Oskari again traveled the mining circuit, sometimes working in hard rock, sometimes in coal mines. Near Sacramento, he married a Finnish girl. In 1900 the Tokoi family returned to Lestijoki, Finland. Oskari fulfilled the miner’s dream by buying a farm and becoming

a merchant. Then he entered “a new career— the obscure and dangerous business of politics.”23 He became active in the Finnish Federation of Labor and the Social Democratic Party and utilized his American labor experience to introduce reforms: an eight-hour day, sickness and accident insurance, maternity assistance, new laws governing domestic and farm workers. In March of 1917, Oskari Tokoi, once coal miner at Carbon, Wyoming, became the first premier of Finland. His moment was brief, as his country plunged into World War I and the Russian Revolution, with civil war on all sides.

1. Daniel Kinnaman, A Little Piece of Wyoming (Kinnaman Publications, Rawlins, Wyo., 1996), 158. 2. Eugene McAuliffe et al., History of the Union Pacific Coal Mines (Colonial Press, Omaha, 1940), 28. McAuliffe became president of the Union Pacific Coal Company in 1923. 3. A. Dudley Gardner and Verla Flores, Forgotten Frontier (Westview Press, Boulder, Colo., 1989), 28. 4. Daily Independent, Laramie, Wyo., 27 February, 1874. 5. McAuliffe, History, 43. 6. Martha Ferguson McKeown, Them Was the Days (Macmillan, New York, 1950), 256. “Lankies” were

Around the same time down in the Little Snake country, a sharpeyed cowboy, W. W. Wilson, was also thinking about oil. In his years of riding through the region from Brown’s Hole to Hahn’s Peak, from Powder Wash to the Sweetwater River, he, like famed geologist David Love a generation later, took note of landforms and geological peculiarities, working out in his own mind the reasons for washes, badlands, springs, oil seeps, mesas, domes. His

miners from Lancashire. One brave miner complained to Superintendent L. R. Meyer that he hadn’t been paid what was coming to him. Meyer carefully went over the ledger and admitted that, indeed, the miner had thirty cents coming, after screenage and leakage were deducted, whereupon Johnson, the miner, replied, “You don’t need to pay me the thirty cents. Just keep it and call it stealage.” 7. Carbon Archive, Hanna Basin Museum, Hanna, Wyo. 8. Carbon County Journal, August 16, 1890, 3. 9. John Edward Grooman manuscript, files of the Saratoga (Wyo.) Sun, c. 1935. 10. Ibid. 11. Carbon County News, July 13, 1878. 12. McKeown, Them Was the Days, 245. 13. Ibid., 255. 14. Mont Hawthorne’s further westward adventures are related in McKeown’s The Trail Led North and Alaska Silver. 15. McKeown, Them Was the Days, 263. 16. Oskari Tokoi, Sisu (Robert Speller and Sons, New York, 1957), 17. 17. Expenses were added up and divided among the fourteen boarders; a fee for cooking and laundry was paid to Mrs. Pelander. 18. Tokoi, Sisu, 40. 19. The league was founded in 1851 to promote temperate and rational living; by 1892 there were four such societies in Wyoming. 20. Tokoi, Sisu, 41. 21. Ibid., 63. 22. Ibid. 23. Ibid., 97.

obituary noted, “Although unschooled, he had an intuitive knowledge of the earth’s structural designs: and through some manner of biologic inheritance was well grounded in elemental geology. His eagle eye could trace faults and structural folds at considerable distance. His accuracy in this respect was uncanny. Skilled geologists often sought his advice.”17 Wilson was making the jump from cow-centric stock raising to the mechanized, big busi-

Opening the Oyster

347

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 348

Uranium Prospecting in the Red Desert dudley gardner

Russia’s entry into the nuclear age spurred uranium prospecting activity in southwest Wyoming. Prospectors armed with Geiger counters spread across the Red Desert. The cold war boom of 1946 –1958 brought a rush of men and vehicles into the desert’s remotest corners. The prospectors shared similar visions with the nineteenth-century gold miners. They wanted to get rich quick. The Geiger counter gave off a telltale clicking sound that indicated uranium-bearing deposits. Somehow, walking over the desert and listening for radiation has the ring of a fool’s quest, but the desert north of Wamsutter and west of Baggs contained millable uranium.

The Red Desert prospector was half adventurer, half entrepreneur. The adventurer drank hard and scrabbled over the land in a World War II surplus jeep; the entrepreneur bought up the adventurer’s claims and formed companies capable of extracting and processing the mineral. One man, Joe Goddard, worked in Craig, Colorado, and looked for uranium in the Red Desert on the weekends. He said: You should have been here in the fifties. That was a hell of a good time. A guy would come into a bar here in Craig and say: “I found a good Geiger reading south of the Haystacks.” Then he would dummy up and never say exactly where, or maybe if he was a big mouth

ness energy extraction business that characterizes Wyoming today. He was an early bridge between the cowboy life and the oilman. He roamed the Red Desert and its margin lands, seeking oil, talking with petroleum geologists who came to depend on his knowledge of local rock formations, drawing information from them, and storing it in his capacious mind. For more than twenty years Wilson watched for promising oil sites. The time given over to the search limited his ability to make a decent living, and only the support of his family allowed him to continue searching. The Powder Wash area, just over the state line on the Colorado side, struck him as extremely promising. But it cost money to drill for oil, and Wilson did not have much. Now approaching middle age, he bought up as much land as he could. He got the lease permits and the services of a Utah oil geologist, and then, lacking money to go farther, took on two partners, one of them reformed ex-outlaw Elza Lay, the other a banker at the Baggs State Bank, 348

he would tell all, but say “that’s my claim.” We would all plan to head for the Haystacks that weekend. The next night [a guy] would come in and say “I found a good place to mine north of [U.S. Highway] 30.” Of course 30 runs all across Wyoming and we’d ply him with drinks to get the location. He got a free drunk, we got a new plan and we’d head north on the weekend. Overall it was good fun and of course we didn’t get rich or I wouldn’t be working here.”1 Joe made his weekly living by the surest way to make money in mining. He “mined the miners.” Joe sold groceries. Since the beginning of mining in the west, the surest way to wealth was to sell supplies, food, shelter, and alcohol

Louis Maupin. Later both partners backed out. Still Wilson persevered, working at odd jobs trying to raise enough money to keep looking for oil. He finally persuaded W. A. Myers, an Ohio Oil Company man, to drill on the promising site. The Rock Springs Rocket excitedly described what happened next: “That huge gasser ‘tore loose’ after midnight one memorable night in October, 1926, with a deafening roar that could be heard a distance of twenty miles as the 67,000,000 feet of steamy-white gas spectacularly belched its noisy message to the oil world, and sent clouds of vapor high above the derrick. . . . It could not be controlled until daylight when the crew braved the steady shower of shales [sic] missiles to close the casing head valve.”18 At that point the site became known as the Florence Wilson No. 1 on the Hiawatha dome in Baxter basin.19 Wiff Wilson, sixty-six years old, was close to the money now, but before it materialized, he was dead in May of 1928 of what

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 349

to the miners. Historian Duane Smith has estimated that for every miner, five people earned a living in surrounding camps and communities.2 Rawlins and Craig supplied Red Desert uranium miners. Baggs provided the beer. Most prospectors were green and had no idea what to look for in the geological formations. When the Geiger counter sounded, they dug. The novice dug, but the more experienced prospector moved around the hot spot area, where radiation was highest, and then isolated the richest deposit. He dug a prospect hole, erected a post, and attached a claim. Then the backhoes came in. The area south of Crooks Gap, for example, became pockmarked with prospectors’ holes. The telltale claims posts stood out like matchsticks in barren areas. The

claims declared who had filed for the property, the location, and the name of the prospector. To be legal, the claims had to be filed in the courthouse in Green River for Sweetwater County, and in Rawlins for Carbon County. Later a permit had to be obtained from the Bureau of Land Management prior to prospecting, since most of the rich deposits were on BLM land. But during the early rush, regulation was scanty, and the stampede to the Red Desert was on.

1. From 1977 to 1985 we conducted hundreds of oral interviews with people born in the late nineteenth and early twentieth centuries. Most of these transcripts and interviews are in the Western

one source named as acute inflammation of the colon.20 In recent years the family has put up a plaque at the site of the first strike.21 The Hiawatha oil field, among others, is still producing. trona The Red Desert contains the world’s largest and purest trona deposits. Trona, better known as soda ash, is used by soap makers, aluminum and glass manufacturers, the paper and chemical industry. In 1938 Mountain Fuel Supply drilled for oil on Union Pacific land west of Green River but instead hit an almost pure (but unwanted) trona deposit. The Westvaco Chlorine Products Company of West Virginia took on the project, adding trona to its basic industrial chemicals list. During World War II the company made barium nitrate for incendiary bombs and chloral for DDT. After the war, it pursued exploration of Wyoming trona deposits. West-

Wyoming Community College library on file in the History Department. Two loose-leaf volumes and one report were generated that contained a variety of interviews. Carolyn Dirkson and A. Dudley Gardner, “Oral Interviews from Northwestern Colorado” (Colorado Northwestern Community College, Rangely, 1979); A. Dudley Gardner, “Oral Histories of Northwestern Colorado” (Western Wyoming Community College, Rock Springs, 1983); A. Dudley Gardner and David Johnson, Department of Environmental Equality 6 (DEQ6), “Abandoned Mine Land Assessment of 26 Mine Sites in Sweetwater County,” vol. 2 (Western Wyoming Community College, Rock Springs, 1989). In 1983 most of the oral interviews dealing with uranium prospecting in the 1950s were lost in a fire. All that survived were the notes. The Joe Goddard interview is from notes taken in Craig, Colorado, in 1977– 1981. 2. Duane Smith, Rocky Mountain Mining Camps (University of Colorado Press, Boulder, 1967).

vaco was not eager to sell when, in 1948, Food Machinery, an innovative and diversity-minded company that had started life as a California fruit-spray pump operation, approached it with an offer to buy. Food Machinery prevailed and added “Chemical” to its name—Food Machinery and Chemical Corporation, today’s international giant FMC. Soda ash was in growing demand, and FMC believed trona could be the source of high-grade soda ash. It took the company several years of research and experimentation to produce trona-based soda ash in large quantities. The Green River refining plant was finished in 1953. Other producers, including giant Solvay Minerals (Belgium) came into Wyoming for the trona, but in the late 1990s China began to give these natural trona producers competition with a synthetic soda ash. The Chinese synthetic soda ash factories are located near seaports. Wyoming plants have difficulty matching their low labor and transport costs.

Opening the Oyster

349

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 350

In the 1970s FMC tried solution mining on trona. A chemical solution was pumped into the trona bed where it dissolved the trona. Later the company went to longwall operations with continuous miner machines. FMC also produced phosphates in Wyoming but closed down the plant in 2004. One worker died in a Solvay trona mine in 1995, and another in 2004 at the OCI Wyoming Big Island Mine and Refinery. uranium Uranium is a mineral, but it is not found in a pure state. It is often the consistency of coarse sand and mixed with earth or rock—the uranium mixture. The amount of uranium in the mix translates into high-grade or low-grade ore. The richest mixtures are the primary minerals — pitchblende and uraninite. In the United States, carnotite, a secondary uranium-bearing mineral, was and is the most common source of uranium, and because uranium is usually present only in minute quantities, complex refining— enriching—is necessary. Geiger counters, scintillation counters, then bulldozers and dump trucks were the uranium prospector’s tools. In the 1920s, radium and vanadium were the important metals. Uranium was an almost worthless by-product of vanadium mining. It was used as a coloring agent in ceramics. In the early 1930s, scientists began to explore nuclear chain reactions after the discovery that splitting atoms resulted in terrific energy discharges. In 1934 Irene Joliet, the daughter of Madame Curie, identified uranium as the best source of atomic energy. By the late ’30s German scientists were working out the details of the fission process. In 1939 Albert Einstein sent a letter to President Roosevelt saying that a uranium bomb could be made. But it was 1942 before the project to develop an atomic bomb started under the Manhattan Engineer District of the Army Corps of Engineers. Suddenly the government realized that most of the country’s uranium came from the Belgian Congo and Canada. It quickly bought up tailings piles from those countries and began an im350

mediate search for a domestic source of the suddenly valuable mineral. The Colorado Plateau had uranium-bearing ores. The bombs that were dropped on Japan came from uranium mined in Uravan, Colorado, later a Superfund site. Gradually prospectors found uranium-bearing soils in other regions, including the Red Desert. Although Wyoming uranium is low-grade (0.04 to 0.23 percent uranium oxide, whereas uranium from some Canadian sites is 20 percent), requiring extensive processing to make “yellowcake,” at the time of the cold war it was desirable. The U.S. government bought uranium, and that supported market kicked off a flood of prospectors. Thousands of people had read of Charlie Steen’s lucky 1952 strike near Moab, Utah, and Vernon Pick’s accidental find, also in Utah. Both men became millionaires. In 1955 the Pumpkin Buttes area of Wyoming was especially hot and attracted a mob of prospectors and claim jumpers, many of them armed.22 In 1956 in the Red Desert north of Rawlins a local restaurant owner, Robert Adams, discovered radiation from uranium-bearing ore near Crooks Gap. He flew over likely country in a small plane equipped with radiation detection and, when the readings were high, would drop a bag of flour to mark the area. He gathered investors’ money to start up a processing mill near historic Split Rock, a landmark for nineteenth-century pioneers. One of his backers was Rawlins doctor C. W. Jeffrey. Adams named the town he built to house his workers “Jeffrey City,” replacing the muchloved Home on the Range post office name. His Western Nuclear Corporation became a major player in uranium production. Supply began to exceed demand, and in 1958 the government announced that it would buy only from processed-ore sources in production before November 24, 1958. Prospecting came to a screeching halt. In the early 1960s the Atomic Energy Commission set up its first “stretch-out” program, designed to keep uranium-producing companies going on the expectation that private businesses would start building licensed nuclear energy plants. Early expectations had been that nuclear power plants would provide a lucrative and continuing market for uranium, but those

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 351

plants were slow to come on line. Gradually the uranium companies sold to the big oil and mining corporations, which could afford to wait for a better day. As the government’s procurement program ended, those yellowcake communities that had failed to diversify faced the bust end of the equation. By the late 1960s Western Nuclear was in trouble. In 1970 it merged with Phelps Dodge and caught a second breath of life, Jeffrey City’s population zooming to 4,000 by 1980. But in the background lurked growing environmental apprehension about radiation’s effect on workers’ health, radioactive material pollution, and disposal problems. The Three Mile Island accident occurred in 1979, and the effect on nuclear plant development and uranium supply was chilling. The bust grew ever deeper. By the late 1980s it was all over, with massive plant layoffs, folding businesses, and outward migration. Jeff City seemed as moribund as Home on the Range. In the late 1990s a slender revival began at the Jackpot Uranium Mine at the base of Green Mountain. Today there is enthusiastic talk among investors, politicians, and energy corporations — even environmentalists — about the benefits of building new nuclear power plants and refurbishing the country’s 103 old reactors. Nearly thirty companies have announced an eagerness to start construction. Although there are extensive uranium-bearing ores in Wyoming, its low-grade character makes it barely profitable to process. Yet in July 2006, SXR Uranium One (a Canada/South Africa company) bought mineral rights and the licensed Jeffrey City uranium mill. A spokesman for the company is enthusiastic about the venture and expects to have the operation running by 2010.23

der River basin and the Red Desert. In the old days the unwanted gas was burned off in great flares. Gas may exist in oil in a dissolved state, or it may rest above the oil as a cap, or be trapped in reservoirs without oil. It may lie near the surface or in deep seclusion. The gas comes from the coal itself, which, because of its countless porous strata, has a large internal surface area. The gas clings to the multitudinous surfaces. The ideal sandwich is gas on top, oil in the middle, water at the bottom. But the ideal is rare. Underground water, which permeates coal beds, traps methane. This water must be drawn off to move the gas to the well. Much of the water is saline, though some is good. Disposal of the water is a problem, as discharging it into streams can cause wildlife habitat damage. To reinject it underground increases production costs, something extraction companies wish to avoid. The first use of natural gas in the United States was in Fredonia, New York, in the early twentieth century. The major problem with gas in the past was the lack of transportation until around 1890, with the invention of leakproof pipeline coupling. Still, the materials were unwieldy and the construction difficult. The gas, considered a nuisance of oil production, was most often flared at the wellhead. In the late 1920s, companies began to lay long-distance pipelines, and by the late 1960s, gas was recognized as an important source of energy. The history of coal-bed methane gas extraction in the Red Desert is an ongoing drama avidly watched by environmentalists, historians, the energy industry, government bureaus, and the citizens of Wyoming.

methane gas Wyoming’s oil resources have been worked for decades, but its extraordinary deposits of coal-bed methane gas are just beginning to be extracted, most of them in coal beds lying beneath the PowOpening the Oyster

351

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 352

Notes 1. Ashley’s free trapper arrangement stimulated the men in the wilderness to great efforts, but it also had drawbacks. The trappers were generally under an obligation to sell their furs to whoever had supplied and provisioned them. As time went on, they were more able to provision themselves and sell their furs to the highest bidder. None of these stratagems was original with Ashley, but the combination of all of them was immense in its influence. 2. John Rolfe Burroughs, Where the Old West Stayed Young (Bonanza Books, 1962), 7. Wheat has the ad appearing in the Missouri Gazette and Public Advertiser. 3. Richard M. Clokey, William H. Ashley: Enterprise and Politics in the Trans-Mississippi West (University of Oklahoma Press, Norman, 1980). 4. Ibid., 102 ff. It should be mentioned that the Arikaras and the Mandans were part of a complex Indian trade network. They correctly saw the increasing white incursions into their territory and business as a major threat. The attack on Ashley was part of a larger revolt against the white invasion that closed the upper Mississippi for some years and funneled aggressive trappers from St. Louis overland to the western trapping regions, inadvertently opening the west. 5. The letter has been lost, and historians rely on James Clyman’s account. 6. The Green River valley region embraced the headwaters of the west’s three great river systems—the Missouri, the Columbia, and the Colorado. The small tributaries and streams that fed the big rivers were the ancient homes of countless beaver, and the rich riparian habitat offered shelter and food to many other animals—furbearers and game. 7. From Burroughs, Where the Old West Stayed Young, 7: “[C]ommencing in 1826, and thereafter until 1840, Brown’s Hole was the scene of the greatest of all the rendezvous staged by the fur companies, the volume of business done there annually exceeding even that transacted at Bent’s Fort on the Arkansas.” Burroughs is mistaken. The rendezvous from 1825 to 1840 took place in many different locations, most of them in what is now Wyoming, with popular sites on Ham’s Fork and Horse Creek. See Fred R. Gowans’s study, Rocky Mountain Rendezvous: A History of the Fur Trade Rendezvous 1825 –1840 (Gibbs-Smith, Layton, Utah, 1985), for the endpaper map and interior photographs of the rendezvous sites. Burroughs’s error may have stemmed from Frank Waters’s highly colored book, The Colorado, in the Rivers of America series (Holt, Rinehart, and Winston, New York, 1946), 177. 8. This vast, raw country was not entirely mapped to a single stan352

dard network until the 1930s, when Works Progress Administration surveyors finished the work. Oliver Morton, Mapping Mars (Fourth Estate, London, 2002), 20. 9. Fergus M. Bordewick, Killing the White Man’s Indian (Doubleday, New York, 1996), 43ff. 10. B. DeVoto, ed., The Journals of Lewis and Clark (Houghton Mifflin, 1953), 484. 11. For example, the heroic paintings of John Clymer, a specialist in sentimental realism, are on the covers of many histories of the fur trade, his big, square-jawed trappers in marked contrast to the background Indians, often depicted with receding chins and open mouths. 12. There was more to the fur trade than beaver. Deerskins, muskrat, fox, wolf, raccoon, lynx, bison, mink, sea otter, and river otter were all worth money. For example, in the 1830s the Jews of eastern Europe honored raccoon caps as the dernier cri, and in the same period the Russian army chose raccoon hats for winter. David Lavender, The Fist in the Wilderness (1964; reprint, University of Nebraska Press, Lincoln, 1998). 13. See Carl P. Russell, Firearms, Traps and Tools of the Mountain Men (Knopf, New York, 1967), 6, for a description of the process of making hats from beaver pelts. 14. Lavender, Fist in the Wilderness, 4. Richard Irving Dodge, The Plains of North America and Their Inhabitants (1876; reprint, Delaware University Press, Newark, 1989), 277–278, saw that these country marriages were to the trappers’ advantages, not useful to the Indians beyond the price the trapper paid to the girl’s father: “A grave trouble to the Indians and one of which I have heard many complaints, is the number of widows and orphans left on their hands by white men. . . . When the trapper was an institution of the plains, he did not consider his outfit complete, unless he had one or more Indian wives. When he went back to the settlement he left them behind. . . . At a council held on the North Platte in 1867, between the Indians and some representation of the Government, one of the Chiefs spoke earnestly and feelingly on this subject. He said that his Tribe was poor, that it could not support the widows and orphans abandoned by white men, and he begged that some special provision might be made by the United States for this class of people.” 15. Local historians Nancy Anderson and Mary Lou Korkow are working on a history of the Carbon-Hanna-North Country region. 16. T. A. Larson, History of Wyoming, 2nd ed. (University of Nebraska Press, Lincoln, 1965, 1978, 1990), 300. 17. Rock Springs Rocket, May 4, 1928, 1. 18. Ibid. 19. A 1962 letter from W. Williams of the Mountain Fuel Supply Com-

Annie Proulx

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 353

pany in Salt Lake City to the Moffat County superintendent of schools in Craig, Colorado, outlines the history of the Hiawatha field, transferred from the Ohio Oil Company to Mountain Fuel Supply Company in 1929. 20. Rock Springs Rocket, May 4, 1928. 21. The plaque reads: “In Honor of / W. W. ‘Wiff’ Wilson / Found Oil Here / in 1909 / thanks Gramps.” The family’s date is earlier than newspaper accounts of the oil strike. The Powder Wash Camp, directly related to Wilson’s explorations and connected to the Hiawatha site by an eightinch pipeline in 1940, has been recorded as a historic site (5MF.1352). A 1988 Affidavit of Cultural Resource Inventories is on file with the State Historic Preservation Office. In 1999 the Environmental Defense Fund named Wyoming as the country’s second-worst refinery polluter in the nation after Texas. Wyoming took first place as the worst releaser of the carcinogen benzene. 22. Harry Kursh, How to Prospect for Uranium (Fawcett, Greenwich, Conn., 1955), 30 –31. 23. Samuel Western, “The Fourth Wave,” High Country News 38, no. 16 (September 4, 2006), 10 –11.

Opening the Oyster

353

Proulx_07_pp.283-354:Layout 1

8/4/08

3:53 PM

Page 354

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 355

26. Red Desert Outlaws Annie Proulx

mericans value pragmatism, competitiveness, and the idea of individualism. By extension we also hold darker values, such as the drive toward domination and conquest, the aggressive determination not to run away from confrontation but to act in self-defense. This stance embodies the legal position and socially pervasive belief that Americans have “no duty to retreat.” An early example is the behavior of W. H. Vanderburgh, a trapper with the Rocky Mountain Fur Company who, with his companions, was set upon by Indians in 1832. Vanderburgh’s horse was shot out from under him, but Warren Angus Ferris, a member of the party, later wrote that “with unexampled firmness, he stepped calmly from the lifeless animal, presented his gun at the advancing foe, and exclaimed ‘boys[,] don’t run.’”1 Richard Maxwell Brown’s examination of the legal history of the English duty to retreat found that “one of the most important transformations in American legal and social history occurred in the nineteenth century when the nation as a whole repudiated the English common-law tradition in favor of the American theme of no duty to retreat: that one was legally justified in standing one’s ground to kill in self-defense.”2 Brown cites a number of state court decisions in the last quarter of the nineteenth century that supported standing one’s ground; frontier attitudes, the turbulent

A

territorial society, election-day partisan rancor, and the concept of masculine bravery all contributed weight to the decisions. The Ohio “true man” case of Erwin v. State in 1876, and the 1877 Indiana “American mind” decision were important.3 In the Ohio case, the state Supreme Court found that a man who killed his son-in-law was, as “a ‘true man,’ ‘not obliged to fly’ from his assailant.”4 The second case, its source a fight in Indiana on election night of 1876, pitted a Democrat and a Republican against each other. While waiting for election returns, the men fought, and Runyan, the Democrat, killed the Republican. Because he had not retreated, he was found guilty of manslaughter. On appeal the Indiana Supreme Court ruled against the earlier judgment, saying that “the tendency of the American mind seems to be very strongly against the enforcement of any rule which requires a person to flee when assailed.”5 In 1921, Supreme Court justice Oliver Wendell Holmes Jr. wrote the crucial opinion on Brown v. United States. Two Texas men had long been enemies. In 1917 both were employed in excavating ground for a new post office building. A quarrel broke out. Brown shot and killed the other man. The district court invoked Brown’s duty to retreat. The case went to the U.S. Supreme Court, where Holmes stated that “the law is a living, growing organism, to be molded by the needs of a changing society.” If some355

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 356

one was under deadly assault, he could stand his ground and even kill in self-defense.6 The belief that a real man had no duty to retreat in confrontational situations was embedded in western frontier life, where the common use of firearms, plenty of gunfighters and outlaws, the lack of official law and order in rough towns frequented mostly by young males where there was nothing to do but drink and fight, created an atmosphere of permissive violence. The telegraph crossed Wyoming in the early 1860s, introduced mass communication, influenced the development of the railroads, and led to a national news service that subjected readers to “an official view of reality.”7 The credulous public absorbed the media content of the day and believed in a wild and woolly west as epitomized in Buffalo Bill, eagerly accepting the most outrageous fabrications as truth. Much so-called reportage was shameless fiction.8 Almost as soon as the last tie spike pounded into the transcontinental railroad, passengers were clamoring to see wild Indians, cowboys on bucking broncos, and shoot-’em-up bad men. Brown translates the mythic categories of hero and villain into two kinds of gunfighters: the down-home “grassroots” cowboyoutlaws invisible in their own time except locally, and the “glorified” gunfighters (such as Wyatt Earp, Billy the Kid) whose deeds still feed thousands of pulp western articles, novels, television shows, and movies. Continuing strong allegiances to the Confederacy or the Union and political bent also differentiated the types as either (dissident) Democrat or (politically conservative) Republican. The conservative (Republican) gunmen tended to support progress and work for men of property and corporate interests. Many of them became sheriffs, deputies, stock inspectors, and Pinkerton men. The (Democratic) dissidents were generally small-town or rural men allied with homesteaders’ interests and resistant to big cattle outfits, the railroad, and, after 1905, the U.S. Forest Service, and eventually the federal government.9 Richard White also sees two types of western violence—personal, one-on-one violence that faded as law and order gained 356

ground; and social violence, sometimes allied with law, government, and political power. For White, “nineteenth century America took great pride in the ability of their political system to defuse, or at least ameliorate, social conflicts. Americans believed in rule of law, not in the rule of force and violence.”10 The showdown, or “walkdown,” as Brown calls it, is, for White, part of the myth of the west “of rugged and armed individualists: gunfighters, outlaws and sheriffs. Myth makes all conflict personal and resolves all conflict with violence.” 11 And Brown, too, in his earlier study of frontier justice, saw local vigilante movements as essentially cooperative, conservative efforts to defend traditional values, investments, and work. The enemies were the criminal elements, especially those threatening community livelihoods. In the raw new towns there was a real fear that low, disorderly elements could take political control. It had happened in other places: “In southern Illinois, in the 1840’s, the ‘flathead’ element of horse thieves, counterfeiters, brigands, slave stealers, and Ohio River-bottom dwellers triggered a violent Regulator reaction. In east Texas, in the late 1830’s, a similar combine of horse thieves, counterfeiters, slave stealers, and ‘land pirates’ incited a countermovement by the Regulators. By 1841, a group of outlaw gangs had virtually taken over the Rock River counties of northern Illinois until challenged by the Regulators in that year.”12 Even closer to home was the almost successful General HalfBreed Bill in Dakota Territory in 1862. The Democrats, realizing that the second-largest (Indians were the largest) ethnic group in Dakota Territory were “half-breeds and trappers of mixed blood,” tried to push through a bill that would permit any of these people who could “read, write, or speak the English language” to become citizens, hoping to harvest their grateful votes. But the specter of “Moccasin Democrats” running the territory roused the Republicans. The bill failed by one vote.13 Brown’s categorization of “incorporation gunfighters” as the politically conservative men who worked for the rich and powerful, and the “resister gunfighters” or dissident social bandits who rebelled against those same power centers, loosely fits the out-

Annie Proulx

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 357

laws active in and around the Red Desert in the 1880s and ’90s. A somewhat different characterization came from outlaw Matt Warner, who divided them into “fighting cowboy outlaws” and criminal killers.14 Various individuals and ephemeral gangs operated in and around the Red Desert in these years. Gustave “Frenchy” Tronquet operated as horse trader and rustler in the early twentieth century. Dutch Charlie Burris of the Parrott gang was snatched off a train in Carbon and lynched in 1878. Big Nose George Parrott himself was later lynched in Rawlins. Dr. John Osborne skinned Parrott’s corpse and had a pair of shoes made from the hide. Elected governor, he allegedly wore the grisly shoes to the inauguration ceremony. Jack Morrow was a drinking, gambling teamster gone bad who robbed travelers and partners. He worked a circuit from Omaha to Denver to Salt Lake City. He showed up in Wyoming in the 1850s, and before the railroad came through, he mined coal around Black Buttes and became something of a local character. When the Union Pacific began laying track through the Red Desert, Jack Morrow got a contract to supply ties and cordwood between Green River and Black Buttes. He bragged about shortchanging the railroad with punky ties and short counts. His name is perpetuated in the northern Red Desert in the Jack Morrow Hills, Jack Morrow Creek, and Jack Morrow Canyon, his early residency apparently overriding his criminal ways. Wyoming outlaws spent much of their time hiding out in rough country where box canyons, thousand-foot cliffs, and blind turns made intimate knowledge of the terrain and water sources essential. The cowboy-outlaw period in Wyoming was brief, roughly from 1875 to 1895. When it was over, the outlaws knew it. Matt Warner and his pals, after the $30,000 Two Bar holdup in 1894 or ’95 stirred up the entire country, tried to make their escape but discovered the old days were finished.15 It wasn’t long before we realized that the whole West had changed, and the days of long flights on horseback was just about over. Every direction we turned we was headed off. All

the bridges, fords, ferries, and mountain passes was held against us. . . . After several wild weeks of chasing and dodging, twisting and turning and hiding, we realized we was bottled up in Wyoming, and the state was alive with posses looking for us. It was hard for us at first to understand this change that had come over the Old West. . . . For a time we couldn’t see that what was behind it was that more railroads, telegraph lines, wagon roads, bridges, farms, cities and settlements was blocking all the old long trails, filling up the old hiding places, and making it easier for the law to spread a dragnet over the whole country.16 It took the gang two years to get out of Wyoming. And a decade later Robert LeRoy Parker (Butch Cassidy) and Harry Longabaugh headed for Bolivia, where something like the rough country and low population of the west they once knew still existed. Some outlaws, as Tom Horn, were unable to leave or to adapt. Utah, with hundreds of small ranches, did not have the rustling problems that afflicted Wyoming and Arizona, the latter both characterized by huge, open-range spreads owned by absentee cattle kings. Huge herds owned by faraway investors powerfully tempted many cowboys with sticky ropes and branding rings. In the 1890s, cattle prices were high and rustling became big business. Men who made a living stealing, rebranding, and selling other men’s cattle stepped up the pace. These men are often called social outlaws because they named their enemies as the absentee cattle kings, the railroad with its vast land holdings and punitive shipping rates, banks owned by fat-cat ranchers and outsider rich men. They were apt to leave gold pieces on some homesteader’s table in return for the borrow of a fresh horse. Poor homesteaders who might themselves never steal a cow or a dollar were deeply pleased when a Matt Warner or a Butch Cassidy took the mighty down. Noble declarations aside, many of these men seem to have been in the game for the sake of comradeship,

Red Desert Outlaws

357

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 358

adventure, and the power-thrill of the face-down. Matt Warner’s report of Bill McCarty’s decision to become an outlaw should be taken with a grain of salt.17 Matt Warner (1864–1938), born Willard Erastus Christianson in Ephraim, Utah, ran away around age fifteen. He found work on a ranch on Diamond Mountain, close to Brown’s Hole and the Wyoming line on the edge of the Red Desert. It was a sixty-mile ride to Rock Springs. The ranch was a rustling operation, and Warner fell into the outlaw life. Perhaps he already had a predisposition to the style, for his brother-in-law was outlaw Tom McCarty. By the time he was in his twenties, Warner had swapped his cattle for horses, “grabbed a ranch site and cabin on the mountain . . . built some corrals, and started a reservoir I never did finish. In less than three years from the time I struck the Diamond Mountain country, I had more than a hundred head of horses of my own I got by broncobusting, trading, breeding, and rustling. . . . About once each season we used to ride into Rock Springs in a whooping bunch and paint the town red.”18 He progressed to holdups and cattle stealing on a large scale. He was friends and temporary partners with most of the regional bad men of the time, including Butch Cassidy and Elza Lay. After years of outlawry, bungled holdups, running and hiding, betrayals, incarcerations, a near lynching, and personal tragedy, Warner went straight in 1900. In Carbon County, Utah, he even served as justice of the peace and deputy sheriff. Robert LeRoy Parker (1866 –1909), a.k.a. Butch Cassidy, was another Utah Mormon boy. Warner and Parker worked together several times. Parker got his outlaw start as a teenager in Utah when he worked for a rancher-outlaw named Mike Cassidy. Cassidy instructed the boy in the art of cow theft, and in some twisted homage, Parker took the man’s name. Cassidy’s “Wild Bunch” often used the Red Desert and its fringe towns for hideouts and blowouts.19 He rustled the big outfits’ cows, sometimes in large batches, robbed banks and trains, and according to Matt Warner’s as-told-to autobiography, when he ran into Cassidy around 1892 after a long separation: 358

[i]t was the same old Butch, except he looked a little older, tougher and harder. . . . Just now he was all het up and righteously indignant about the Johnson County War . . . a war of extermination that had just been declared by the big cattle kings and their cowboys up there against the little cattle rustlers. They was shooting on sight, hanging or dragging to death at the end of lariats anybody caught rustling or stealing cattle or horses. . . . Butch and me and Tom was all for the preservation of the ancient and honorable calling of cattle rustling, and the perpetuation of the Old West against all those new-fangled things that was changing the Old West, like farms and cities, railroads and telegraphs, and whittling down the wild, free, open country and making prisoners, captives, and sissies out of free-riding cowboys and now we could fight back.20 Cassidy, reputedly easygoing and good company, worked out certain rules that let him see himself as a man of honor. He never killed a man in the commission of a holdup. When he and his companions worked as cowboys, they did not steal cattle from the home outfits, which made them popular hires. They generally targeted the big spreads for their depredations. And Cassidy’s Robin Hood behavior after a train or bank holdup, by scattering money for small services such as a dinner of fried chicken at some lonely homestead, made him a welcome visitor.21 In their own way the Wild Bunch contributed to the local economy. Outlaws hid out with friendly ranchers on the edges of the Red Desert in Lander and Brown’s Hole.22 In the desert proper they had secret hideouts, including upper Powder Springs, a little northeast of old Sulphur Springs (lower Powder Springs) on the Cherokee Trail just a few miles from the Colorado border. Dripping Springs, a few miles farther east, was also an outlaw hideout, according to Matt Warner.23 Wyoming cowboy L. H. “Doc” Chivington reinforces that in his autobiography, “Last Guard”: “From Powder Springs we led circle to Dripping Rock

Annie Proulx

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 359

Springs which is fine water. As we had been drinking such bad water, we all went to the spring for a drink. We were sitting around tanking up when someone noticed some cans under a flat rock. We started investigating and pulled out a large can of flour, several cans of hominy, pork and beans, and other canned food which had no doubt been cached there when the famous Powder Springs outlaws were in this country years before.”24 David Darlington and Josh Bodyfelt have unearthed a quitclaim deed “issued by Matt Warner to H. H. Metcalf for the sale of eighty acres . . . encompassing Lower Powder Spring.”25 Warner undoubtedly knew the surrounding country extremely well. An old stone cabin, reputedly used by Cassidy, is still halfstanding near upper Powder Springs, rugged country of hoodoos, juniper, and ancient cairns. Ute, Shoshoni, trapper, explorer, outlaw, stockman, and settler pasts all come together in this remote corner of the state. Beneath overhanging ledges are hundreds of human and animal charcoal figures drawn by the oldest inhabitants. A few miles away a curious stone “throne” stands on a steep hill overlooking a flat arena. Stone hunting cairns mark lookout points. A large, relatively level meadow is partially enclosed by seven miles of fence mostly fashioned from slashed and bent-over junipers in the style of English hedgerow makers. Sections of the fence are missing, and one can guess that sheepherders or cowboys used them for firewood, as rusted cans and charcoal fire signs show up near the fence gaps.26 The Cassidy cabin is hidden near a gully a few hundred yards below the meadow. Although Cassidy’s Wyoming State Penitentiary mug shot for the only term he served shows a rumpled, cross-looking teddy bear with a black eye, the reader should make no mistake: Butch Cassidy was a hard man. He had a reputation as being rough on horses. The long rides, the tough companions, the bold robberies, the extremely rugged and demanding country he and his followers inhabited, made adamantine men with quick reflexes and no regrets. In one of the few Cassidy letters believed to be authentic, he wrote from Santa Cruz, Bolivia, in November 1907: “We would

of left here long ago, but we had a little trouble with the old mule. Ingersoll hobbled her and tied her to a tree and wore a nice green pole out on her, but I didn’t think he had done a good job so I worked a little while with rocks. Between us we broke her jaw and we have been feeding her on mush ever since, but she can eat a little now and we will leave in a few days.”27 The Wild Bunch broke up before the hideouts at upper Powder Springs and Dripping Rock Springs had neighbors. In 1905, L. H. “Doc” Chivington set up a homestead just a few miles away, and cowboy-entrepreneur W. W.“Wiff” Wilson was nosing around the region looking for oil seeps. Elza Lay (1868 –1934), born William Ellsworth Lay in MacArthur, Ohio, came west and got a job driving a horse car in Denver. He thought he had killed a drunk passenger after an altercation and fled northwest to Brown’s Hole and the surrounding Utah-Wyoming country. In 1896, when Butch Cassidy finished his only prison term, he began a serious recruitment of men for his gang, soon referred to as “the Wild Bunch.” Elza Lay, well known in the Red Desert border regions, became Cassidy’s lieutenant. Lay married into a ranch family on Diamond Mountain the same year. The marriage was brief. The couple divorced after wife Maude discovered Lay’s part in several robberies, and before daughter Marvell Lay was born in 1897. According to Matt Warner, at one time Lay visited him and suggested they rob a Jewish peddler together. Cassidy and Lay pulled off the Pleasant Valley Coal Company payroll heist at Castle Gate, Utah, in April 1897. After dividing the money at their Robber’s Roost hideout, the gang descended on Baggs, Wyoming, for a whoop-up forever engraved in Baggs’s history. By 1899 Lay and Cassidy were in New Mexico, where, to Cassidy’s disapproval, Lay took part in the second Folsom train robbery on July 11, 1899. Lay was wounded after a posse traced them to their hideout, and soon afterwards arrested, charged, and sentenced to life. His sentence was commuted for good behavior, and he was released in January 1906. He came back to Baggs and went straight for the rest of his life. He, like cowboy Wiff Wilson,

Red Desert Outlaws

359

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 360

became interested in oil, and according to one source, it was Lay, not Wiff Wilson, who discovered what later became the Hiawatha oil field.28 Nothing came of the oil venture for Lay, and he ended up in California as watermaster for the Imperial Valley irrigation system. Bob Meldrum was an alcoholic and professional man-hunter who lived for a time in Dixon and worked as a harness maker for Little Snake storekeeper-entrepreneur Charley Perkins. Meldrum was characterized by John Rolfe Burroughs as “intolerant, contumacious and captious. . . . His fits of anger bordered on insanity.”29 He supposedly shot fourteen men dead in Colorado mining camps before coming to Wyoming. Around 1908 the [Little] Snake River Cattlemen’s Association hired him to control an outbreak of cattle rustling. The sheriffs of Routt County in Colorado and Carbon County in Wyoming moved to make Meldrum a deputy, but the public outcry forced the Routt County sheriff to give up the idea. The Wyoming sheriff did make him a deputy. By 1910 the cattle rustling situation had cooled off, whether because of Meldrum’s attentions is unknown. The town of Baggs appointed Meldrum city marshal, an error in character judgment that cost two men their lives. One of them was Chick Bowen, a popular young cowboy who made the mistake of standing on the sidewalk and “whooping and yelling.” 30 Tried, released under bond, and tried again, Meldrum was sentenced to the state pen for five to seven years, but apparently spent the time as a trustee working on saddles and harness on William Daley’s ranch west of Rawlins. When his time was up, he opened a harness shop in Walcott under the patronage of Charles Vivion’s Leo Sheep Company and eventually disappeared. He fit Matt Warner’s category of criminal killer better than cowboy outlaw. Big cattlemen supported him when the law pressed too sharply on him. He knew Tom Horn, perhaps more a business acquaintance than friend.31 Horn was a hired gunman who came to Wyoming in 1892 and worked for various ranch owners who wished passionately to halt rustler depredations. Horn worked for the Two Bar, part of the Swan empire then under the management of John Clay Jr. 360

Some believe Horn was involved in the Johnson County War at the request of Clay. He allegedly told Wyoming governor W. A. Richards, himself a ranch owner with a rustler problem, that “whenever everything else fails, I have a system which never does.”32 Horn’s “system” consisted first of riding into a rustling area and letting himself be seen. This was sometimes enough to encourage rustlers to leave. If they stayed, he would pin an “orelse” note on their doors. If they stuck it out, he found a good position and shot the man. He dispatched two rustlers in the Hanna-Rawlins area before concentrating on the area around Iron Mountain. He was based on the Iron Mountain ranch, owned at the time by John C. Coble and Frank Bosler. After killing Willie Nichols, the son of a rancher in the Iron Mountain locale, Horn was taken, tried, and hung in Cheyenne, November 1903. Law and order had replaced hired killers in Wyoming. The wild badlands and tiny water seeps still awaken thoughts of men gone wrong, hiding in the most remote places they knew, alas, remote no longer.

Annie Proulx

Proulx_08_pp.355-376:Layout 1

8/4/08

3:54 PM

Page 361

Notes 1. Lyman C. Pedersen Jr., Warren Angus Ferris, and LeRoy R. Hafen, The Mountain Men and the Fur Trade of the Far West (1965; reprint, Arthur C. Clarke, Spokane, 2001), 2:147. Vanderburgh was shot dead. 2. Richard Maxwell Brown, No Duty to Retreat: Violence and Values in American History and Society (Oxford University Press, New York and Oxford, 1991). 3. Ibid., 8 –26. 4. Ibid., 9 –10. 5. Ibid., 17. It was not only in frontier settlements that this attitude prevailed. David J. Langum found four examples of self-defense in impromptu trials on the emigrant trails; see his “Pioneer Justice on the Overland Trails,” Western Historical Quarterly 5, no. 4 (October 1974): 428, 431, 435– 436. 6. Brown, No Duty, 34. Holmes also remarked that “detached reflection cannot be demanded in the presence of an uplifted knife.” 7. Menahem Blondheim, News over the Wires: The Telegraph and the Flow of Public Information in America, 1844 –1897 (Harvard University Press, Cambridge, Mass., 1994), 2. 8. J. Anthony Lukas, Big Trouble: A Murder in a Small Western Town Sets Off a Struggle for the Soul of America (Touchstone, New York, 1997), 632 – 686. Today the inflated and vividly imagined misdeeds of outlaws are still relished, rewritten, and retold, a kind of persistent fakelore that fuels tourism and television’s junk history programs. A murky field known as “outlaw history” leaves the serious historian, after encountering numerous repeated errors, loath to trust any of the books. In the case of Butch Cassidy, for example, most useful are Anne Meadows, Digging Up Butch and Sundance (St. Martin’s Press, New York, 1994), and F. Bruce Lamb’s The Wild Bunch: A Selected Critical Annotated Bibliography of the Literature (High Plains Publishing, Worland, Wyo., 1993), wherein most often appear the words “did not,” “no evidence,” “full of errors,” “inaccurate,” “mislabeled,” and “untrue.” To be fair, Lamb also makes a few small errors, such as putting Brown’s Hole in Wyoming, not Colorado. 9. Brown, No Duty, 60, comments: “By the 1860s, man-to-man gunfighting was an established practice in the West. Down to 1900, there were thousands of Western gunfighters. As grassroots gunfighters, the vast majority of them earned neither national nor regional fame and were ignored by myth makers.” Traditional values, including clan loyalty and vengeance, dominated behavior. If you called a man a liar or a son of a bitch, you could expect gunplay. All the same, a Dr. Charles B. Penrose

wrote Owen Wister in 1892 that “during the past two months ‘son of a bitch’ has been a favorite expression in this country. Wyoming is in the son of a bitch stage of her civilization and could not get along without it any more than she could without a lariat and a branding iron.” T. A. Larson, History of Wyoming, 2nd ed. (University of Nebraska Press, Lincoln, 1978), 307. 10. Richard White, “It’s Your Misfortune and None of My Own”: A New History of the American West (University of Oklahoma Press, Norman, 1991), 328. 11. Ibid. 12. Richard Maxwell Brown, Strain of Violence: Historical Studies of American Violence and Vigilantism (Oxford University Press, New York, 1975), 106. 13. Howard Roberts Lamar, Dakota Territory 1861–1889 (Yale University Press, New Haven, Conn., 1956), 87– 88. 14. Matt Warner and Murray E. King, The Last of the Bandit Riders (Bonanza Books, New York, 1940), 224. Bob Meldrum, Harry Tracey, and Tom Horn would fit into the criminal killer category. 15. The Two Bar was one of the Swan Cattle Company’s numerous ranches. Ranchers did not like to make such losses public, for fear it would give other holdup artists ideas. John Clay, manager for the Swan holdings, avoids any mention of such a holdup in his memoir. Big ranch cattle sale money could top bank robbery takes. Mari Sandoz, The Cattlemen (University of Nebraska Press, Lincoln, 1958), 494, credits the Two Bar with “160,000 cattle and 200 riders.” She wrote that the ranch put on the first rodeo in the early 1880s, during which Butch Cassidy, who briefly rode for the brand, gave a pistol-shooting exhibition. 16. Warner and King, Last of the Bandit Riders, 229. 17. From Warner, in Bandit Riders, 177: “At that time he [Bill McCarty] was bitter and desperate about losing everything he owned on account of the depression. He was drinking hard to drown his troubles and was going to pieces. He blamed Wall Street, the big corporations, the railroad companies, the politicians, and the government for wiping him out.” 18. Ibid., 42. 19. According to Lamb, in The Wild Bunch, 11–12, the men did not come together as a gang until after the Montpelier, Idaho, bank robbery of 1896 and the Castle Gate, Utah, payroll robbery of 1897. Neither Matt Warner nor the McCartys were part of the Wild Bunch. 20. Warner and King, Bandit Riders, 213–214. 21. Cassidy was so well liked that decades after his 1909 death in the Patagonian shootout, Wyoming people swore they had seen him in var-

Red Desert Outlaws

361

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 362

ious towns, probably after encounters with the imposter William T. Phillips. 22. Butch Cassidy sometimes hid out with a rancher named Brown, whose spread was not far from Lander. Many outlaws stayed with Charley Crouse or Matt Warner, both on Diamond Mountain. 23. In Bandit Riders, 236 –237, Warner describes this hideout, which he used, as “in the middle of the big desert that laps over southern Wyoming and northern Colorado. At the head of a canyon water drips from a sandstone cliff into a catch basin. The only ranch in that part of the country was thirty miles away and owned by a man named Tyson. The place was a natural rock fort. . . . We camped under some piñons near the water quite snug and comfortable. . . . We hadn’t any more than got rested up when we thought this place was so lonesome we couldn’t stand it.” 24. L. H. Chivington, “Last Guard”: Memories of an Old Time Cowboy from 1901 to 1923 (unpublished manuscript, Museum of Northwest Colorado, Craig, n.d.). E. W. Leggett, a Baggs resident who was interviewed by the Colorado WPA in the 1930s, remarked: “The Cassidy headquarters was not in Powder Springs proper but at lower Drippin’ Water, a tributary of Powder Springs.” Civil Works Administration, Moffat County, Colorado, 1933–1934 interviews, pamphlet 356, State Historical Society of Colorado, 122. 25. David Darlington and Josh Bodyfelt, “Outlaws and Horse Corrals,” Wyoming Archeologist 43, no. 2 (Fall 1999): 80. “The instrument was filed on October 12, 1885, and the date of record is December 19, 1897. It was witnessed by E. H. Rife, a rancher in the Cold Springs Mountain area west of Powder Wash. The Quit Claim deed is on file at the Sweetwater County Courthouse, Green River, Wyoming” (ibid.). 26. David Darlington, D. Murcray, and Josh Bodyfelt have studied the history of what they refer to as the “drift fence” over the last decade and, based on tree-ring growth counts for junipers in the fence, have concluded it was built by ranchers and/or outlaws in the late nineteenth or early twentieth centuries to hold horse herds. Yet juniper has varying and unreliable growth rates in its different limbs. Despite their research it would have been helpful if they had consulted Bruce Lamb’s bibliography of the Wild Bunch, which points up some of the major errors in the literature. For example, the report that a gang of seventy-five outlaws from the Hole-in-the-Wall country near Kaycee came to upper Powder Springs is erroneous. And to call the fence a “drift fence“ seems misleading; the drift fence originated in the Texas panhandle, not as an enclosure, but as a very long east-west fence, sometimes with wings, to keep cattle from drifting south in blizzards.

362

27. Meadows, Digging Up Butch and Sundance, 99. 28. Pearl Baker, The Wild Bunch at Robber’s Roost (Abelard-Schuman, New York, 1971), 178. 29. John Rolfe Burroughs, “Bob Meldrum, Killer for Hire,” unidentified publication, folder B-M482-rd, American Heritage Center, University of Wyoming, Laramie, n.d. 30. Ibid. 31. Chip Carlson, Tom Horn (Beartooth Corral, Cheyenne, 1991), 59 –65. 32. Ibid., 31.

Annie Proulx

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 363

27. History of Conservation Efforts in the Red Desert Mac Blewer

y some accounts, the Red Desert is the largest unprotected and undeveloped high-elevation desert left in the United States. A Wyoming Outdoor Council brochure, however, notes, “Despite this distinction, the area is the focus of multinational oil, gas and mining corporations. According to the Bureau of Land Management, this pressure will continue to grow, with the industry hoping to turn southwestern Wyoming into the major natural gas producing region in the United States by 2015.”1 True to form, the Red Desert is fast becoming one of the hot spots of natural gas development in the United States. After Wyoming’s Powder River basin and the upper Green River valley, the Red Desert is quickly succumbing to drilling and road building associated with mineral development. Approximately 90 percent of southwestern Wyoming’s public land is available for oil and gas drilling and development. Numerous oil, gas, and coal-bed methane exploratory wells, compressor stations, coal mines, and trona mines dot this region, linked together by a growing web of service roads, giant overhead power lines, and pipelines. Such development fragments wildlife habitat and can severely disrupt animal behavior and migration. However, despite this reality, many of the desert’s open spaces and wild lands remain relatively untouched. The Jack Morrow Hills Study Area, of the South Pass–Steamboat Mountain coun-

B

try in the northeastern Red Desert, and Adobe Town, of the southern desert, are two such areas that must be conserved. Indeed, for over one hundred years, citizens in Wyoming and around the country have attempted to permanently protect the Red Desert. Today groups are urging Congress to protect this special place as a national conservation area, calling for the buyout or trade of controversial mineral leases while allowing for responsible grazing, hunting, and recreation to continue in the unique region. Despite decades of support and numerous attempts to preserve it, the Red Desert has yet to be protected. Decisions about the future of the region are in the hands of the Bureau of Land Management (blm). The first recorded effort to protect the Red Desert was an 1898 proposal by a Lander hunter, Dr. Frank Dunham, to designate a winter game preserve stretching from the Oregon Buttes–Steamboat Mountain area all the way to Yellowstone National Park. His proposal, which was championed by the local sporting group, the League of Associated Sportsmen (las), was outlined in Recreation Magazine: We were discussing the question of establishing a winter game park, and concluded to put the l.a.s. in possession of some facts, and to call its attention to a stretch of country 363

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 364

that is practically of no value to any one, but which furnishes winter range for more elk and antelope than all the remaining . . . portion of Wyoming. Look at a map of this state, and in the northern part of Sweetwater county you will find a tract of country marked ‘the red desert.’ We conclude that by setting off a portion of country between the twelfth guide meridian on the East, and the thirteenth guide meridian on the West, the fifth standard parallel (North) on the South, and the South boundary of Fremont County on the North, enough of this territory would be taken in to furnish a winter range for the elk, and other large game for all time. . . . I have seen large bands of elk, in November, making their pilgrimage to this red desert country. . . . We are thoroughly aroused to the necessity of immediate action if we could afford adequate protection for our game, and if any interested sportsmen desires further light on the subject Mr. Dodge and I will gladly furnish it. This is a most excellent proposition, and the officers of the l.a.s. will undoubtedly present a bill to Congress, at its next session, asking that the Park limits be extended so as to include the portion of the Red Desert indicated on the above map.2 Although this proposal never got off the ground, it started a robust movement alive and well over a century later. By the early 1900s, sheep and cattle grazing in the Red Desert had escalated to such a point that the ecosystem was in serious jeopardy in the “Land Where God Ran Out of Mountains,” as the Arapahos called it. Warnings such as this one followed: “In 1918 the Secretary of the Interior reported that the Red Desert–Great Divide Basin country was so badly overgrazed that it would be lost if measures were not taken to spare it.”3 Relief did not come to the fragile Red Desert ecosystem until the 1934 Taylor Grazing Act was passed. In 1935, Wyoming governor Leslie Miller and the Wyoming delegation crafted a bill for the U.S. Congress that would have created the Western Trails Na364

tional Park, stretching all the way from Missouri to Washington State. Along the 2,000-mile proposed route, such scenic and historical attractions would have included Fort Laramie, Fort Caspar, Independence Rock, Devil’s Gate, and South Pass. As reported in the May 9, 1935, issue of Wyoming State Journal: A bill drawn by the Wyoming commission recently named by Governor Miller had been forwarded to Congressman Greever. It provides that “the regions immediately contiguous to these old trails comprise areas of great historical consequence embracing pioneer farms, sites related to early fur trade history, Indian battle grounds, a historical monument, overland stage and pony express stations, outstanding landmarks and prehistoric excavations and deposits of unusual geological and archeological significance.4 Although the legislation did not succeed, it did elevate the importance of the pioneer trails and of the South Pass historic landscape. In 1943, Jackson Hole National Monument was created through executive order by President Franklin D. Roosevelt in cooperation with John D. Rockefeller, who had for several years been buying up land in and around the area. The executive order, and the resulting creation of the monument was akin to betrayal, according to some such as former Wyoming governor and U.S. senator Cliff Hansen. Mr. Hansen in an interview stated, “A number of boys were in the service then . . . and almost every serviceman from Jackson Hole felt that he’d a knife stuck in his back. They felt that they were away from home fighting to protect their country, and while they were away from home, their homes had been taken away from them.”5 According to one particularly acerbic journalist, Westbrook Pegler, the executive order was along “the general lines of Adolf Hitler’s seizure of Austria. They [President Roosevelt and Interior Secretary Ickes] anschlussed a tract of 221,610 acres for [Interior Secretary] Icke’s domain. . . . A man who might be likened to the Nazi governor of Poland.”6

Mac Blewer

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 365

At the time, the social divisions within Wyoming ran deep. Martin Murie, a biologist and writer and the son of pioneering Jackson Hole conservationists Mardie and Olaus Murie, remembers those times well: All kinds of verbal hell broke loose. I remember the day because my dad, Olaus, and I, returning from some fieldwork in the northern part of the valley, stopped for mail and met a rancher, Buckenworth, storming out of the post office. He stopped and told Olaus the news; That man in the White House etc. . . . Olaus took it calmly, made no commitments, didn’t argue. That was his way. He was on the other side of the fence and Buck knew it, but he had to spout just the same and it is also possible that Buck thought the “outrageous” act by that man in the White House might just possibly push Olaus into rage on the ranchers’ side. But no.7 Later on, Mr. Murie recounts that things grew even more interesting: So, time for spring roundup and driving of cattle to summer ranges. The ranchers made plans, strapped on their sixshooters and rounded up a bunch of Herefords for a challenge drive across National Monument land. Wallace Beery happened to be in town, probably on movie business, but I’m not sure, and they roped him into the scheme, everybody knowing ahead of time that cattle leases had been carefully grandfathered into the National Monument charter, including the right to cross federal land with the herds. But the national media liked the defiance, since most of that media was overtly or covertly in the anti-FDR camp. So they attended with cameras and reportage: armed ranchers driving cattle, rangers standing beside their pickups watching the show. The Herefords were duly delivered to north and east

ranges. Nothing happened. Beery and the ranchers had a little pistol practice on the ride home, looking for targets such as ground squirrels, bottles, cans. I don’t know if anybody hit anything. Those six guns were sometimes carried by ranchers and cowboys, in saddle gear or on a belt, “for coyotes.” (For show, say I. There has never been a walk-down on the streets of Jackson and I know of only one case where a pistol was drawn and fired at a live and hostile man. By Fred Deyo, the game warden, but that’s another story.) Robert Hansen and I, classmates in high school, sat in his car one day to discuss the situation. We decided that even if our dads were on opposite sides in the fight, the two of us could still be friends. The storm passed, but the ideology remained. Wyoming is still mired in that antigovernment pose. It is a pose. I don’t think it comes from a philosophical anarchist doctrine; it derives directly from old territorial days when ranchers, some of the really big ones being investors from England, considered themselves above the law and acted accordingly. They used “the commons,” i.e. public lands held by the national government, as their own domain and bitched like hell when the national government asserted its rights. I have to add that corporate interests of all sorts, especially in the “wrecreation” and energy industries, still assert themselves in powerful ways on public lands. That too is another, very big, story. The moral, if there is one here, is that sometimes a very rich man can do good. Just think of what we would see and endure now if developer sprawl had taken over those lands north of Jackson!8 Another crucial point came in 1950 when Grand Teton National Park was created through the expansion of Jackson Hole National Monument. Enraged legislators crafted an amendment to the Antiquities Act, which read, “The Act of February 26, 1929 (45 Stat. 1314), and any other provisions of law heretofore specif-

History of Conservation Efforts in the Red Desert

365

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 366

ically applicable to such present park or monument, are hereby repealed: Provided, That no further extension or establishment of national parks or monuments of national parks or monuments in Wyoming may be undertaken except by express authorization of the Congress.” The conservation community swallowed the compromise in order to allow Teton’s expansion. To this day Wyoming is the only state (except for Alaska) that does not allow for the creation or expansion of national monuments or parks through executive order. Congressional action is required for such a measure to become law. In 1963 the National Park Service conducted a survey of the Red Desert. Attending were Mr. Herbert Pownall of the University of Wyoming; Dr. David Love; Mr. Bryan Harry, assistant park naturalist of Grand Teton National Park; and several other Park Service and Bureau of Land Management staff. Mr. Tom Dustin of Fort Wayne, Indiana, was invited but could not be present. The expedition visited Boar’s Tusk, the Sands, Steamboat Mountain, Oregon Buttes, Continental Peak, and the Honeycombs. As one person reported: While the dunes around Essex Mountain were interesting and are quite extensive, we feel that Great Sand Dunes National Monument contains more impressive and massive dunes similarly deposited against an updraft producing mountain slope. . . . The Honeycomb Buttes area was the most impressive seen during the visit. The coloring is very attractive and varied, as are the erosional forms. However, we question its significance and degree of difference from Bryce Canyon, which is also eroded from a similar number of the Wasatch Formation, or from Badlands, Theodore Roosevelt, Cedar Breaks, Capital Reef, and perhaps other similar areas already in the National Park System.9 Tom Dustin, conservation chair of the Indiana Division of the Izaak Walton League and frequent desert traveler, said: “What the Park Service failed to take into account was the value of the Red 366

Desert in its entirety. It is like no place on earth! There are few places in North America where you can get so many different landscapes, so many different values coming together in one area.” By the early 1960s, fencing had become a major issue in the desert country. During one survey the blm found some 6,000 fences covering a distance of 13,000 miles, many of them illegally built. The Bar X fence proposal near Oregon Buttes in the South Pass Historic Landscape became a huge issue of contention between local conservationists and Mr. Leonard Hay of the prominent ranching and minerals family of southwestern Wyoming. Although the proposed fence would cut across big-game migration corridors, the proposal was still approved. Mr. Hay would keep his private property in the northern desert open to hunters and anglers and was known to keep his cattle numbers at low levels to match the forage and drought conditions. A highly charismatic and influential figure, Mr. Hay was respected by many, although his views on oil and gas development and the role of the rancher would be questioned by some.10 In 1968, Wyoming Outdoor Council founder Tom Bell recommended that much of the desert be designated a North American antelope range. The proposal would have reduced livestock grazing and fencing. Mr. Bell traveled to Washington, D.C., with Cora rancher Gale Noble and presented their proposal to then interior secretary Stewart Udall and the Wyoming delegation. Udall reacted favorably. However, Senator Gale McGee told them that the proposal was not “politically feasible.”11 Around this time, serious discussions were held between conservationists, hunters, resource management officials, and ranchers regarding the reintroduction of bison into the Great Divide basin. The last bison had been shot near Oregon Buttes in 1898.12 Although ranchers had attempted to manage several herds of bison in the 1950s, the beasts proved unmanageable and dangerous.13 In 1973 the Wyoming Game and Fish Department made the first effort to introduce the concept to the blm, in a memo from Mr. Max Long, deputy game warden in Rock Springs:

Mac Blewer

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 367

Today there is a great amount of feeling being expressed by environmentalists and sportsmen of not only this state, but of many other regions throughout the entire United States to attempt to have this great animal once more restored his natural range. This animal once restored and protected could be utilized for hunting. . . . In conclusion gentlemen, I would like that you consider this as a formal letter asking that in the Great Divide Area known as the Red Desert you consider the allotting of aum’s in the amount sufficient to sustain 1000 head of bison winter-count. Said increase of animals to be hunted as a game animal of Wyoming. I believe that you would agree that you not only would be giving consideration to the residents of this state, but rendering a just service to the people of the United States.14 The proposal, like later proposals to restore bighorn sheep to the Sweetwater Rocks area of the northern desert, would be met with vocal opposition from the ranching community. But since the Red Desert was, by some accounts, the largest open, unprotected area in the lower forty-eight states, bison proposals would be advanced off and on for the next few decades. In the 1970s the Park Service surveyed Adobe Town, saying: “The greatest natural value of this area is it [is] still a ‘howling wilderness.’”15 In the survey, the Park Service concluded that the area was worthy of national natural landmark status due to its significant ecological and geological values. In the winter of 1983–1984 a wildlife tragedy that garnered national media attention shone the spotlight on the Red Desert and the greater issue of illegal fencing. Hundreds of pronghorn perished along Red Rim in the southern desert, trapped within the tripartite confines of a snowdrifted rancher’s fence, the I-80 corridor, and the Union Pacific Railroad right-of-way. “I remember walking the tracks and seeing carload after carload of antelope that had been hit. It was a gut-wrenching experience,” recalls Dr. Steve Torbit, director of the National Wildlife Federation’s

Rocky Mountain Natural Resource Center, who was working with the Wyoming Game and Fish Department at the time.16 Eventually, after pressure from the National Wildlife Federation, the Wyoming Wildlife Federation, hunters and game officials, and furious attention from the national media, Governor Ed Herschler successfully forced the property owner to take the fence down. Later the Wyoming Game and Fish Department bought Taylor Lawrence’s land, 24,000 crucial acres of some of the best pronghorn habitat in North America. Union Pacific gave the National Wildlife Federation the coal lease underlying the property.17 In the early 1970s, renowned geologist Dr. David Love, conservationists Dick Randall and Tom Dustin, and others championed the unique characteristics of the desert and the need to protect some of it from development. Several wilderness proposals were drawn up for parts of the desert, notably for Steamboat Mountain, Buffalo Hump, the Sands, and other areas. In the late 1970s and 1980s, Wilderness Society Wyoming representative Bart Koehler, Wyoming Wildlife Federation president Tom Dougherty, Atlantic City resident and outfitter John Mionczynski, Rock Springs resident Craig Thompson, Jack Peugh, Bruce Hamilton, and others focused the conservation community’s attention on the desert and its swiftly changing nature due to the oil boom, the construction of the Jim Bridger Power Plant, the increase of trona and coal mining, the continued erection of lethal barbed-wire fences, and other development issues. Recalls Bart Koehler, director of the Wilderness Society’s Wilderness Support Center, “Dick Randall and I would lead citizen trips in the desert. He would do the nature stuff. I would do the political stuff. And we would stop at Steamboat, Bush Rim, and then the Honeycombs. . . . Of course, the development of the Bridger Power Plant would change the nature of everything. It put a lot more pressure on the region.”18 After the Federal Land Policy and Management Act was passed in 1976, the blm was directed to conduct wilderness inventories on its lands. It would make recommendations for twelve

History of Conservation Efforts in the Red Desert

367

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 368

wilderness study areas throughout the Red Desert country in 1991: Buffalo Hump, Sand Dunes, Oregon Buttes, Whitehorse Creek, Sweetwater Canyon, Honeycomb Buttes, Alkali Draw, South Pinnacles, Red Lake, Alkali Basin/East Sand Dunes, Red Creek Badlands, and Adobe Town. In 1992 the blm issued the draft Green River Resource Area Management Plan. Hunters and conservationists stood up for the desert and once again reiterated the need to protect parts of it from large-scale development. Two years later, in response to blm’s wilderness study area proposal, the Wyoming Wilderness Association, a coalition of Wyoming conservation groups, presented to blm a “Citizens’ Wilderness Proposal for Wyoming blm Lands.” The proposal recommended that many of these wilderness study areas be enlarged and designated wilderness, while allowing for limited development to occur in other areas of the desert.19 In 1997 the blm issued the final Green River Resource Management Plan that deferred decisions on fluid minerals and locatable minerals for the 88,000-acre core area surrounding Steamboat Mountain. It was decided that an amendment to the Green River RMP would be drafted for this core area and the entire 623,000-acre Jack Morrow Hills Study Area. The document would be called the Jack Morrow Hills Coordinated Activity Plan. In February of 1998 the blm issued a notice of intent to prepare a coordinated activity plan for the Jack Morrow Hills Area. That June the blm offered several oil and gas leases inside the study area. A coalition of conservation groups successfully protested. In July the blm halted all oil and gas leasing in the study area, putting in place a temporary moratorium on mineral leasing, oil and gas development, and most surface disturbance until the planning process was complete. The blm’s 2000 draft environmental impact statement calling for development of the northern Red Desert came out in June. The hearings in Lander and Rock Springs were dominated by conservationists, tribal members, and hunters who testified regarding the importance of the area to big game, scenery, spiritu368

ality, cultural heritage, and open spaces. Nearly 13,000 comments were received by the agency, 93 percent of them favoring the Citizens’ Red Desert Protection Alternative, which had been drafted by conservationists favoring the trade or buyout of mineral leases in the northern desert. At the time, this was the largest number of comments ever received by the Bureau of Land Management for a Wyoming project.20 Interior secretary Bruce Babbitt visited the Red Desert in November of 2000, proclaiming the need for the blm to restart the planning process for the Jack Morrow Hills Study Area. After a tour of South Pass Historic Landscape with Tom Bell and Wes Martel, of the Eastern Shoshone Business Council, he stated: I want to call a time-out in this process and take a step back. . . . I say with great confidence that surely we want the [blm’s environmental analysis] to begin with the presumption that we can protect this landscape. . . . I am going to instruct the blm to do that with a conservation alternative as a preferred alternative. . . . I’m not here to sign a piece of paper making a designation. But I think that the dialogue and discussion have not gone far enough. The draft EIS, in my judgment, does not match the quality, the beauty and the expanse of the Red Desert. I want to bring this out into the full sunshine of Wyoming, the West and the nation.21 The secretary’s visit was met by angry protesters at Riverton airport, including Fremont county commissioner and People for Wyoming leader Crosby Allen, and with similarly outraged statements from Governor Jim Geringer and Senators Thomas and Enzi. On December 22, 2000, Secretary Babbitt issued a memorandum to the director of the blm to choose a conservation alternative as a preferred alternative in a “supplemental draft plan.”

Mac Blewer

Having reviewed this Plan with my staff, and having visited the area, I am greatly impressed by the unique and out-

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 369

standing natural resources contained in the planning area. With one of our Nation’s largest unfenced areas outside of Alaska, its big game populations are among the largest and healthiest in the lower 48. It contains one of the most diverse and numerous concentrations of raptors anywhere. Significant cultural resources, including remnants of the Oregon and Mormon Pioneer Trails and the mining camps of South Pass, only add to the area’s allure. . . . Any decision to protect the outstanding natural resources of the Red Desert must be accomplished in a manner that protects the valid existing rights of these mineral owners. To the extent it is consistent with our paramount concern for protecting the natural resources of the planning area, some additional leasing might be allowed, but the presence of finite mineral resources should not deprive future generations of the natural and aesthetic wonders of the Great Divide Basin. A final decision as to how this area should be protected will necessarily be made by a future Administration. Nonetheless, it is my responsibility to place the blm on a track that helps insure that, after a full opportunity for public participation, an appropriation decision will be made to protect this unique area and its outstanding resources. To that end, I ask that you direct the Wyoming blm office to propose the conservation alternative as its preferred alternative in the supplemental draft EIS that the Solicitor has determined should be prepared.22 Shortly thereafter, Governor Geringer and a coalition of mining, oil, and gas companies and ranchers petitioned the Department of the Interior to stop the administration from implementing the conservation alternative as a preferred alternative. The National Wildlife Federation and thirty ranchers, outfitters, businesses, and conservation groups responded to the petitions by formally asking the new interior secretary, Gale Norton, that they also be included in any deliberations regarding the northern Red Desert.

On November 13, 2001, the Wyoming blm state director, Al Pierson, submitted a formal memo to the Rock Springs Field Manager, reiterating the sentiments of Ms. Nina Hatfield, acting director of the blm, in a personal memo written in September of 2000 in addition to the memos of former secretary Babbitt and then interior solicitor John Leshy. Mr. Pierson stated: You will note that both memoranda make it very clear that our effort will be a supplement to the existing draft and that we are to complete the analysis of the comments received on the original draft and then analysis of a full range of alternatives, including a Conservation Alternative and a Preservation Alternative. It is also very clear from the memoranda that the “Preferred Alternative” will be selected only after thorough analysis of a full range of alternatives. This process is consistent with blm‘s NEPA [National Environmental Policy Act] regulations.23 Although it was determined by many legal analysts that Secretary Babbitt had not overstepped his boundaries when it came to providing guidance to the blm, the agency determined to take a softer, less controversial route to the Jack Morrow Hills planning process by starting the process anew with a supplemental draft plan with a wide range of alternatives, including a conservation alternative and a preservation alternative. A few months earlier, in 2001, the Wilderness Society, the Wyoming Chapter of the Sierra Club, the Wyoming Outdoor Council, and Biodiversity Associates announced plans to win national conservation area designation for the Red Desert. The proposal called for the buyout or trade of mineral leases in the Jack Morrow Hills Study Area while “preserving key wildlife habitat, cultural and historic resources and the unique Red Desert landscape.” The proposal left the door open to include other scenic and scientifically important landscapes in the desert, such as the Big Sandy area and Adobe Town. As Bart Koehler, executive director of the Wilderness Society’s

History of Conservation Efforts in the Red Desert

369

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 370

Wilderness Support Center, exhorted supporters at woc’s annual meeting in June 2001, “It’s time to saddle up, start on a long, hard ride and work together to keep Wyoming a wild place, today, tomorrow and until the last sunset blazes across our skies.”24 In February 2003 the Supplemental Draft Plan for the Jack Morrow Hills was released, calling for increased oil and gas development in the northern desert. The plan was met by criticism from tribal leaders, conservationists, and hunters. In May of 2003 the Eastern Shoshone and Northern Arapaho business councils, in separate resolutions, called for the blm to halt any future oil and gas development in the Jack Morrow Hills Study Area and asked that greater protections be put in place for Native American – respected places within the desert such as Steamboat, the Sands, Boar’s Tusk, White Mountain, Oregon Buttes, and the Indian Gap Trail:

ure,” yet non-Indian sites such as the California and Oregon National Historic Trails, the Mormon Pioneer Trail, and the Pony Express Trail, are all given a three mile viewshed protection zone! Non-native historical sites such as the Crookston Ranch, the Tri-Territory Marker and the South Pass Historic Landscape were all withdrawn from surface mineral development. And who makes these standards? The blm! The petroglyphs, medicine wheels, etc. . . . have been there long before the Oregon Trail and are more spiritual/ancient, yet, the non-Indian sites receive stronger standards than a Native American sacred site. Why? With the involvement of both tribes, we can set our own standards, we just need to speak up and actually be there during the planning process. We need to be at the table, and not on the menu! It’s stated in President Clinton’s Executive Order #13007 that cultural sites are to be defined by the tribes, not the blm! These sacred sites where the spirits of our people live are now threatened with oil and gas development. These developments open up new roads that are linked together by a web of power lines, pipelines and more roads. . . . Our ancestors have for centuries revered the natural springs and associated them with cleansing and fertility. The Red Desert has countless springs that provide life for all of creation. Those springs are now in jeopardy! Our ancestors from the past have been protecting these sacred sites for centuries, this has always been our homeland security, and we owe it to them to continue protecting what we can!25

To date, we believe that the Bureau of Land Management (blm) has produced a plan that fails to protect Native American holy sites and parts of the Jack Morrow Hills Study Area that were used by our ancestors for hunting, medicine gathering and spiritual purposes. The Young Warriors Society, a youth empowerment group for the Eastern Shoshone and Northern Arapaho tribes on the Wind River Reservation, organized several slide shows and a run through the Red Desert to raise public awareness of the issue. In a May 22, 2003, editorial in the Wind River News Martin Blackburn of the Young Warriors Society stated: Not only do the wildlife cherish this area, but our ancestors did as well . . . Our ancestors also say that the sand dunes are alive and the spirits protect certain areas as well . . . The blm proposes a buffer zone (where there is to be no surface disturbance, like oil rigs, roads or mining within that distance from the site) of 100 feet for Native American sacred sites such as the Indian Gap Trail, as a “protection meas370

The plan garnered over 69,000 comments, the largest number of comments ever received by the blm for a single Wyoming planning project. Approximately 98 percent of the comments were in favor of the Citizens’ Wildlife and Wildlands Protection Alternative, a similar alternative to the Red Desert Protection Al-

Mac Blewer

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 371

ternative, but written with more input from the ranching community. At hearings in Rock Springs and Lander in April of 2003, 80 percent of the citizens spoke in favor of greater protections for the area. Among the citizens testifying was Peter Absolon of the National Outdoor Leadership School (nols). Absolon, former assistant director of the nols Rocky Mountain branch, stated, “We believe it is important for the blm to preserve [the area]. Some places should just be left alone. The Jack Morrow Hills is one of those places. We strongly recommend that the blm draft a conservation alternative that does not include oil and gas drilling.”26 Mark Soldierwolf, an Arapaho Indian, commented, “We’ve lost a lot of land that is so dear to us. Not only Native Americans but all of us. We need to sit down and say, that’s enough.”27 In 2004 the blm issued the Final Environmental Impact Statement and Record of Decision for the Jack Morrow Hills Study Area. It was met with over one thousand protest letters from citizens across the nation for its many failings. Wyoming governor Dave Freudenthal again voiced his frustration with the blm, especially with the vague nature of the plan, the unknown pace and scope that the development would take, and the potential impacts that the plan would have on big game. In the same year, the draft Great Divide Resource Management Plan (also known as the Rawlins Resource Management Plan) was released, calling for the development of large stretches of the Great Divide country. The plan would direct the agency on the management of 4.7 million acres of land, including the eastern and southern half of the desert, Adobe Town, the Ferris Mountains, and other wildlands for the next twenty years. The plan called for the development of approximately 9,000 wells, over 2,500 miles of roads, utility lines, and associated developments.28 The Friends of the Red Desert submitted the Western Heritage Alternative, a balanced alternative calling for phased oil and gas development in some areas and the protection of wildernessquality lands. Hearings were swamped by pro-desert protection

advocates, and the blm once again received tens of thousands of comments in favor of protecting the Great Divide country. Longtime Casper resident and self-proclaimed Red Desert rat Carol Long stated, “I can get closer to God in the desert than anywhere else. I pray that the blm will take the actions suggested by concerned citizens who love the magical place called the Greater Red Desert.”29 In July of 2006 the blm issued the Final Jack Morrow Hills Coordinated Activity Plan and Record of Decision, calling for the development of the northern desert. It is unclear how much development would be allowed under the guidance of the plan, and with what pace. After a meeting with conservationists in Washington, D.C., shortly after the plan was released regarding designation as a national conservation area, Senator Craig Thomas issued a press release that read: [The senator says:] “After a nine year public process, many protections have been put forward that I believe adhere to multiple-use strategy for public lands. . . . However, the public needs to remain engaged and the implementation of this plan needs time to play out. . . . These protections are the direct result of public involvement, opportunities for protection in the environmental process and provide for multiple use of a portion of those lands within Jack Morrow Hills. “It may be appropriate as time goes by to provide for wildlife corridors, to protect some of the open space areas, to ensure that wildlife can make the appropriate movement within the Hills.” Thomas said that there may also be a possibility of a National Conservation Area designation at some point in the future, however, at the time being it seems appropriate to give the public time to evaluate the plan before any changes.30

History of Conservation Efforts in the Red Desert

371

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 372

The blm will undoubtedly approach the Jack Morrow Hills plan implementation with much caution. The agency has already been put on notice regarding the sensitive nature of gold exploration near Oregon Buttes in the South Pass Historic Landscape. And the agency is gingerly handling the Great Divide planning process. With 2,000 wells slated for Atlantic Rim, and the continental divide, Fontanelle, Moxa Arch, and Jonah Field oil and gas drilling projects booming, the desert is changing. But citizens are paying attention. And they are standing up for the desert, as they have for years. As Bart Koehler recently said: It is not necessarily the dedication of one or two people over time. It is clearly the devotion and dedication of many steadfast citizens over a long span of time. But there’s something else that’s very special at play out there where those wild winds blow. It is a phenomenon that has tugged at the hearts of many people over several lifetimes. I call it being “thunderstruck.” At least that’s what happened to me the first time I set foot near Honeycomb Buttes. If you look over time, whatever magic is out there in the desert, in those wild open spaces, it has drawn people to speak up to protect it for centuries. It is the magic of the place. Looking out over the Red Desert from Oregon Buttes, it represents the essence of so much of what is good and real about Wyoming. Sooner or later we will get lasting protections for this Wild Heart of the West.31

372

Notes 1. Wyoming Outdoor Council, Wyoming’s Red Desert: Wild Heart of the West (brochure, 2001). 2. Frank Dunham, “A Winter Game Preserve,” Recreation (1898): 271–272. 3. Tom Bell, “Wyoming’s Red Desert” (unpublished article, 1988). 4. “Plan Great National Park to Memorialize and Protect Scenes along Oregon Trail,” Wyoming State Journal, May 9, 1935. 5. kued, The Jackson Hole Story (one-hour documentary, 2004), http://www.kued.org. 6. Robert W. Righter, Crucible for Conservation: The Struggle for Grand Teton National Park, edited by R. Righter (Colorado University Associated Press, 1982), 115–116. 7. Martin Murie, personal correspondence with the author, October 13, 2006. 8. Ibid. 9. David Hieb, regional chief of boundary studies for the National Park Service, to nps regional director, “Field Study of Boar’s Tusk–Sand Dunes–Red Desert Area, Wyoming,” January 21, 1963. 10. Leonard Hay, interview with the author, Rock Springs, Wyo., summer 2002. 11. Tom Bell, interview with the author, Lander, Wyo., August 1999. 12. Tom Bell, interview with the author, Lander, Wyo., September 1998. 13. Leonard Hay, interview with the author, Rock Springs, Wyo., summer 2002. See Gary P. Beauvais, “Vertebrate Wildlife of the Red Desert,” this volume. 14. Max Long, director of blm, Rock Springs District Office, Rock Springs, Wyo., to wildlife professionals, regarding buffalo restoration in the Great Divide Basin of Wyoming, March 24, 1971. An aum (animal unit month) is the basis for grazing privilege fees and permits allocated by the Bureau of Land Management (blm) and the usda Forest Service. 15. Biodiversity Conservation Alliance, Adobe Town Briefing Book (Laramie, Wyo., 2004), http://www.voiceforthewild.org/general/pdfs/ Adobe_Town.pdf. 16. Dr. Steve Torbit, interview with the author, Boulder, Colo., August 2006. 17. Tom Dougherty, interview with the author, Loveland, Colo., August 2006. 18. Bart Koehler, interview with the author, Durango, Colo., August 2006.

Mac Blewer

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 373

19. Liz Howell, interview with the author, Sheridan, Wyo., June 1999. 20. Andy Tenney, interview with the author, Rock Springs, Wyo., 2001. 21. Jeff Gearino, “Babbitt: Red Desert Plan Flawed,” Casper Star Tribune, November 16, 2000. 22. Bruce Babbitt, secretary of the interior, Washington, D.C., to director of blm, “Jack Morrow Hills Coordinated Activity Plan,” December 22, 2000. 23. Al Pierson, blm state director for Wyoming, to the Rock Springs field manager, “Guidance for Preparation of Jack Morrow Hills Supplemental (jmh) eis,” November 13, 2001. 24. Mac Blewer, “Red Desert Dreams,” Frontline (newsletter of the Wyoming Outdoor Council), Fall 2001. 25. Martin Blackburn, “Young Warriors Seeking Protection for Red Desert,” Wind River News, May 22, 2003. 26. Susan Gray Gose, “Desert Lovers Blast blm’s Plan,” Lander (Wyoming) Journal, April 13, 2003. 27. Ibid. 28. Erik Molvar, interview with the author, Laramie, 2004. 29. Wyoming Outdoor Council, “blm Schedules Hearings for Public Comment on Rawlins RMP,” press release, January 15, 2005. 30. Office of U.S. Senator Craig Thomas, “Thomas: Too Early to Consider National Conservation Area Designation for Jack Morrow Hills,” press release, July 31, 2006. 31. Bart Koehler interview, August 2006.

History of Conservation Efforts in the Red Desert

373

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 374

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 375

contributors

Nancy Anderson has been wandering the history of Carbon County, Wyoming, for nearly fifty years. Her companions and mentors were Vera Oldman of the Grand Encampment Museum and Muriel Kitching of the Hanna Basin Museum.

Laura Fertig studied the Red Desert of Wyoming for her master’s degree in botany from the University of Wyoming in 1997. She currently resides in Kanab, Utah, and is an animal care giver/naturalist/dilettante.

Dr. Gary Beauvais is the director of the Wyoming Natural Diversity Database, a service and research unit of the University of Wyoming that develops information on the state’s rarest plants and animals. He has studied Rocky Mountain wildlife for sixteen years, focusing primarily on habitat use, conservation biology, and biogeography of vertebrates.

Walter Fertig was formerly the state botanist with the Wyoming Natural Diversity Database of the University of Wyoming. He currently resides in Kanab, Utah, and is a botanical consultant.

Mac Blewer served as outreach coordinator for the Wyoming Outdoor Council for six years and continues to work on the Red Desert campaign as a Wyoming Outdoor Council (woc) board member and a volunteer for Friends of the Red Desert. He has been engaged in wildlife conservation and public lands protection campaigns for over twenty years. He is an amateur naturalist, writer, activist, and guide. Charles Ferguson is a structural geologist residing in Centennial, Wyoming, with degrees from Kansas University, New Mexico Tech, and the University of Calgary. He is also an adjunct at the University of Arizona, a research geologist with the Arizona Geological Survey, and a research associate with the New Mexico Bureau of Geology.

Dudley Gardner, Ph.D., has written numerous articles and several books on Wyoming. A historical archeologist by training, he pursues his primary goal of writing histories about people who left no written record, but rather left details of their lives in the form of rock art and artifacts scattered across the desert. George P. Jones, as the vegetation ecologist at the Wyoming Natural Diversity Database since 1988, has conducted studies in the plains, basins, and mountains throughout Wyoming. He lives in Laramie with his wife, Connie Wilbert, and their two children. Jeffrey Lockwood is a professor of natural sciences and humanities who shares his time between the Department of Philosophy and the mfa program in creative writing at the University of Wyoming. He has spent twenty years as an insect ecologist, wandering and wondering his way across the empty spaces that fill Wyoming.

375

Proulx_08_pp.355-376:Layout 1

8/4/08

3:55 PM

Page 376

Andrea Orabona has worked for the Wyoming Game and Fish Department as the nongame bird biologist since 1992, managing populations of over 250 species of birds that are not hunted. She is also president of the Red Desert Audubon Society, which encompasses a membership base from Dubois to Evanston, Wyoming. Annie Proulx received an m.a. in history from Sir George Williams University (now Concordia) in Montreal. After completing coursework and orals, she abandoned her doctoral dissertation and switched to the writing of fiction based on the French Annales approach to historical research. The study of history continues to be a major interest.

Russel L. Tanner is founding director of Kyak Marook Heritage Research, a small archaeology, anthropology, and history research organization based in Rock Springs, Wyoming. He served for twenty years as a U.S. Bureau of Land Management cultural resource specialist in southwestern Wyoming and holds degrees in anthropology and American studies from the University of Wyoming. Craig Thompson teaches engineering and environmental science at Western Wyoming Community College. He has been wandering and studying the Red Desert since 1960, when he moved there with his family. He was educated at the college where he teaches and at Colorado College and Stanford University.

Tom Rea is the author of Bone Wars: The Excavation and Celebrity of Andrew Carnegie’s Dinosaur (University of Pittsburgh Press, 2001) and Devil’s Gate: Owning the Land, Owning the Story (University of Oklahoma Press, 2006). He lives in Casper and writes mostly about the past. Jack States lives in Lander, Wyoming, and is an adjunct professor of botany at the University of Wyoming. He moved back to the state after twenty-five years as a biology professor and administrator at Northern Arizona University in Flagstaff. His books include Mushrooms and Truffles of the Southwest (University of Arizona Press, 1990) and Wildflowers of Wyoming (with Diantha States, Mountain Press, 2004).

376

Contributors

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Page 377

index

Italic page numbers refer to photos, figures, tables, and maps.

A Absaroka Range, 218 Absaroka volcanic field, 91, 98n61, 109 Absolon, Peter, 371 Acari (mites), 192 –193 Accipiter gentilis (northern goshawk), 146, 153 Achnatherum hymenoides (Indian ricegrass), 206, 209, 219, 220, 241 Acoma Indians, 185n4 Acrididae, 194 Actitis macularius (spotted sandpiper), 173, 174, 176, 179 Adams, Charles (Indian agent), 290, 291n22 Adams, Charles F., Jr., 276 Adams, John Quincy, 185n1 Adams, Madeleine Baker, 315n13 Adams, Robert, 350 Adams-Onís Treaty of 1819, 181, 185n1 Adobe Town, 8, 10; conservation of, 81, 363, 367, 368, 369, 371; as feature, 305; geology of, 91 Adobe Town Rim, 91 Aeoloplides, 190 Afar Triangle, 96n8

Agelaius phoeniceus (red-winged blackbird), 172, 173 agriculture: and environmental change, 123; and maize, 233, 234; and Meeker, 287, 288; and plant diversity, 219; and wheat cultivation, 273 Agropyron cristatum (crested wheatgrass), 218 Ahlbrandt, T. S., 100nn107–108, 210 Albertine ranch, 318 alfalfa (Medicago sativa), 218 algae blooms, 108 –109, 117 Algonquin speakers, 249 Alkali Basin/East Sand Dunes, 368 Alkali Draw, 368 alkali flats, 125, 306 alkaline clays, 221 alkaline salts, 108 alkali sacaton (Sporobolus airoides), 221 alkali sagebrush (Artemisia arbuscula ssp. longiloba), 202, 204, 208, 219 Alkali Springs, 117 Allen, Crosby, 368 alluvial deposits, 92 alluvial plains, 108, 110

Altiplano-Puna Plateau, 88 Ambystoma tigrinum (tiger salamander), 147, 148 Amelanchier alnifolia (serviceberry), 121, 212, 240 American avocet (Recurvirostra americana), 173, 174, 179 American badger (Taxidea taxus), 145, 148, 149, 152, 207 American coot (Fulica americana), 173, 179 American cordillera, 88 – 89, 98n67 American crow (Corvus brachyrhynchos), 173, 179 American goldfinch (Carduelis tristis), 145, 153, 173, 177 American kestrel (Falco sparverius), 172, 173, 176 “American mind” decision, 355 American Mustang and Burro Association, 166n74 American redstart (Setophaga ruticilla), 172, 173 American robin (Turdus migratorius), 172, 173, 177 American white pelican (Pelecanus erythrorhynchos), 172, 173, 179

American wigeon (Anas americana), 173, 179 American Wild Horse Preservation Campaign, 166n74 Ammotragus lervia (Barbary sheep), 145, 160 amphibians, 144, 148, 152 Amphispiza belli (sage sparrow), 147, 149, 153, 161, 171, 173, 177 Anas acuta (northern pintail), 171, 173, 176 Anas americana (American wigeon), 173, 179 Anas clypeata (northern shoveler), 173, 176, 179 Anas crecca (green-winged teal), 171, 173, 176 Anas cyanoptera (cinnamon teal), 171, 173 Anas discors (blue-winged teal), 171, 173 Anas platyrhynchos (mallard), 171, 173, 176, 179 Anas strepera (gadwall), 173, 176 Ancestral Rockies, 90 Anderson, Harold, 331 Anderson, Nancy, 345–347 Antarctica, 97n32, 114 antelope. See pronghorn 377

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

antelope bitterbrush (Purshia tridentata), 211, 212 Antelope Hills, 208, 212 Antelope ranch, 319, 323 Anthicidae, 195 Anthomyzidae, 196 Antilocapra americana (pronghorn). See pronghorn Antiquities Act, 365–366 antlike flower beetles, 195 Apaches, 183, 186n19, 249 aphids, 191, 193, 195, 196 Aquila chrysaetos (golden eagle), 146, 148, 149, 171, 173, 175, 179 Arachnida, 193 Araneae (spiders), 193 Arapahos, 249, 275, 364 Archaen seas, 90 –91 archaeological sites: and Fort Bridger, 124, 272, 273–274, 279n4; preservation of, 79. See also petroglyphs; projectile points; rock art Archaic period, 232, 235n3, 250 Archean-aged sialic nuclei, 89, 97n29 Archean granites, 89, 97n31 Archean orogenies, 89 –90 Arctodus simus (short-faced bear), 147, 155 Ardea herodias (great blue heron), 172, 173 Areas of Critical Environmental Concern, 223 aridity: and biological soil crusts, 227; and environmental change, 121, 122 –123, 122, 129; and fire, 207; and plant diversity, 218; and ranching, 317; and wildlife, 144, 145, 150; and winds, 144, 210 Arikaras, 340, 352n4 378

Page 378

arroyos: and Ashley, 126; and beavers, 124–125, 128 –129, 134n71, 156; and environmental change, 124–125, 128, 129; and Stansbury, 134n66; and Twain, 127 Artemisia, 201 Artemisia arbuscula ssp. arbuscula (little sagebrush), 202 Artemisia arbuscula ssp. longiloba (alkali sagebrush), 202, 204, 208, 219 Artemisia bigelovii (Bigelow sagebrush), 202 Artemisia cana (silver sagebrush), 203, 204, 211, 219 Artemisia cana ssp. cana (basin silver sagebrush), 202 Artemisia cana ssp. cana (plains silver sagebrush), 211 Artemisia cana ssp. viscidula (mountain silver sagebrush), 202, 210, 211 Artemisia nova (black sagebrush), 201, 202, 204, 208, 219, 223 Artemisia pygmaea (pygmy sagebrush), 202 Artemisia rigida (scabland sagebrush), 202 Artemisia rothrockii (timberline sagebrush), 202 Artemisia tridentata (big sagebrush), 201, 203, 205, 220, 222 Artemisia tridentata ssp. spiciformis (subalpine big sagebrush), 202 Artemisia tridentata ssp. tridentata (basin big sagebrush), 153, 202, 203, 206, 209 –211, 212, 219, 223 Artemisia tridentata ssp. vaseyana (mountain big sagebrush), 202, 203, 210, 211–212, 213, 219

Artemisia tridentata ssp. wyomingensis (Wyoming big sagebrush). See Wyoming big sagebrush Artemisia tripartita ssp. rupicola (Wyoming three-tip sagebrush), 202, 208 Artemisia tripartita ssp. tripartita (three-tip sagebrush), 202, 203 Arthropoda, 192 Ashley, William Henry, 126 –127, 133nn49 –50, 134n54, 339 – 343, 352n4 Ashley National Forest, 14 Asian markets, and transcontinental railroad, 303n2 Asio flammeus (short-eared owl), 173, 175 Asio otus (long-eared owl), 173, 175 aspen (Populus tremuloides): and birds, 171, 172, 174, 175, 176, 177; and insects, 196; and wildlife, 222 Astor, John Jacob, 339 Athabaskan groups, 234 Athapaskans, 249, 250 Athene cunicularia (burrowing owl), 145, 149, 152, 173, 175, 177 Atlantic City, 275 Atlantic orogenies, 90 Atlantic Rim, 163, 372; view from, 44 Atlas Range, 88 Atomic Energy Commission, 350 Atriplex confertifolia (shadscale saltbush), 206, 220, 326 –327n26 Atriplex gardneri (Gardner saltbush), 206, 220 Avonlea projectile points, 234 Aythya valisineria (canvasback), 173, 179

Index

B Babbitt, Bruce, 368, 369 Babcock, Orville E., 264n27 baculites, 138 badlands, 91, 220 Baggs, 33 Baggs, George, 313 Baggs, Maggie, 313 Baker, Jim, 313 Baltic rush ( Juncus balticus), 221 Bannocks, 234, 248, 271, 272 Barbary sheep (Ammotragus lervia), 145, 160 Barnes, W. C., 222, 225n24 Barnes site, 234 Barney, John Rodney, 332, 336 –337n9 Barnhart ranch, 318 Barr Ell Seven ranch, 332 Bartleson, John, 239, 240, 242 –243nn3– 4 Bar X fence proposal, 366 basin big sagebrush (Artemisia tridentata ssp. tridentata), 153, 202, 203, 206, 209 –211, 212, 219, 223 basin silver sagebrush (Artemisia cana ssp. cana), 202 basin wildrye (Elymus cinereus), 207, 210, 211, 219 Basque sheepherders, 326n8 Bassett, Anne, 321 Bassett Gang, 321 bats, 147, 152 Battle Mountain, 22 Battle of the Rosebud, 286, 291n19 Baxter shale, 138 Beadle, J. H., 294–296, 296n6 Bean, Leonard, 331 Bear Creek, 178 Beard, Chris, 139 –140

Proulx_09_pp.377-400:Layout 1

8/4/08

Bear Lake, 126 Bear Mountain, 95 Bear River, 93, 94, 126, 128, 239, 240 Bear River Divide, 241 Bear River Mountains, 241 Beaton ranch, 319 Beauvais, Gary, 80, 191 beaver (Castor canadensis): and arroyo cutting, 124–125, 128 – 129, 134n71, 156; extirpation of, 156, 165n35, 222; protection of, 158; status in Red Desert, 145; and trappers, 124, 125–128, 129, 134n71, 156, 305–306, 340, 352n6; and water, 99n68, 152, 156 Beebe, Ruth, 326n10 Beery, Wallace, 365 bees, 195 Beetle, Alan, 210 Bell, Tom, 366, 368 Bennet, Amos, 312 benthos, 109 bentonite, 307 Bering land bridge, 140 Bernard, Hi, 321 Berry, Wendell, 198 Bessie Bottom site, 234 Bidahochi, Lake, 92, 93–94, 99nn73,90 Bidwell, John, 128, 239, 240, 242 –243nn3– 4 Bigelow sagebrush (Artemisia bigelovii), 202 Bighorn Basin, 212 Big Horn Range, 218 bighorn sheep (Ovis canadensis): and disease transmission, 157, 160 –161; extirpation of, 222; hunting of, 267; and mountain environment, 153–154;

3:51 PM

Page 379

population of, 158; status in Red Desert, 145 big sagebrush (Artemisia tridentata), 201, 203, 205, 220, 222 Big Sandy Area, 369 Big Sandy River, 95, 113, 114, 116, 125, 128 Billy the Kid, 356 Biodiversity Alliance, 79 Biodiversity Associates, 369 biogeographic provinces, 218, 224n7 biographical rock art, 245, 246 biological diversity: of birds, 193; of insects, 77, 189 –198; and landscape features, 152; and mountains, 153–154; of plants, 194, 217, 218 –220, 221, 222 –223, 224n3, 225n24 biological soil crusts, 227–228 birch, 221 birds: in autumn, 178 –179; diversity of, 193; migration of, 172, 174, 175, 177, 178, 179; peripheral vertebrates, 144; and sagebrush, 153, 171, 176 –177, 179; in spring, 172, 174; in summer, 175–178; surveys of, 171; and water, 152 –153, 171, 172, 174, 177, 179, 318; widespread vertebrates, 148; in winter, 179 –180. See also specific birds Bishop Conglomerate, 98n64 Bishop Mountain, 254, 263n5 bison (Bos bison): and Archaic period, 232; and environmental change, 121, 125, 127; extirpation of, 157, 158, 163, 166n50, 222; and horses, 155; hunting of, 231, 234, 267, 287; and Late Prehistoric period, 233, 234; and Paleoindian period, 231;

protection of, 158; reintroduction of, 366 –367; and rock art, 246, 249; status in Red Desert, 145; trails of, 343; and Union Pacific Railroad, 79 Bison Basin, 140 Bitter Creek: basin of, 77; Beadle on, 295–296; and environmental change, 127, 129, 134n66; geology of, 99n87; hydrology of, 116, 117, 119n41; and Killpecker Creek, 85; and maps, 254; and railroad grading, 294–295 Black, Mary Louisa, 241–242 black-billed magpie (Pica hudsonia), 173, 179 Blackburn, Martin, 370 Black Buttes, 163 Black Buttes Gray, 233 black-capped chickadee (Poecile atricapillus), 173, 177, 179 Blackfoot, 184, 234, 249 black-footed ferret (Mustela nigripes), 145, 149, 160, 163 black greasewood (Sarcobatus vermiculatus), 206, 207, 210, 220, 224n18 Black Rock, 94, 95, 100n101 Black Rock Flat, 85, 94 Black Rock petroglyph panel, 245–248, 249 blacks, 285, 291n10 black sagebrush (Artemisia nova), 201, 202, 204, 208, 219, 223 bladed paths, 340 Blewer, Mac, 80 – 81 Blue Gap, 93 bluegrass (Poa), 219 blue-gray gnatcatcher (Polioptila caerulea), 173, 178 blue-green algae, 227

Index

Blue Point site, 131nn2,10, 231 Blue Rim, 91 blue-winged teal (Anas discors), 171, 173 Boardman, W. P., 289 Boar’s Tusk, 17, 175, 305, 366, 370 Boar’s Tusk rock art site, 246, 248 Boar’s Tusk Ware, 233 bobcat (Lynx rufus), 145 Bodyfelt, Josh, 359, 362n26 Bolton’s ranch, 319 Bone Draw, 113, 114 Bonneville, Benjamin Louis Eulalie de, 344 bonytail (Gila elegans), 145, 159 Bos bison (bison). See bison Bosler, Frank, 360 Bos taurus (cattle). See cattle bow and arrow, 233 Bowen, Chick, 360 box elders, 126 Boy Butte, 305 Brachylagus idahoensis (pygmy rabbit), 144, 147, 149, 152, 153, 206, 208 Braconidae, 195 Bradley, H., 91 Brannan Reservoir, 175 Branta canadensis (Canada goose), 173, 179 Breeding Bird Survey (BBS), 171 Brewer’s blackbird (Euphagus cyanocephalus), 172, 173 Brewer’s sparrow (Spizella breweri), 145, 161, 171, 173, 177 Bridger, Jim: and Ashley, 339; and Bridger Pass, 255, 263–264n14; on the Evans route, 256; and Fort Bridger, 124, 271, 272, 273–274; and Fraeb’s Post, 311; and Native Americans, 271; and Quien 379

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Hornet Mountain, 263n5; and sheep, 325n1; spoilation claim with War Department, 273, 279n2; and Brigham Young, 272 Bridger Coal Mine, 344 Bridger Pass, 126, 254, 255, 264n14, 274 Brisbin, James S., 317 Brislawn, Robert, 335 broad-tailed hummingbird (Selasphorus platycercus), 172, 173, 174 Brockmeyer, John, 311, 313 Bromley, Isaac, 280n35 Bromus inermis (smooth broome), 218 Bromus tectorum (cheatgrass), 157, 159, 161, 212 –213, 219, 222 Brontosaurus giganteus, 137 Brooks, Merle, 308 Brooks, Patricia, 308 Brooks ranch, 319 Brown, John Lowery, 257, 264n23 Brown, Richard Maxwell, 355– 357, 361n9 Browne, Percy, 293 Brown’s Hill, 26 Brown’s Hole, 233, 234, 341, 352n7 Browns Park, 98n65 Brown v. United States (1921), 355– 356 brucellosis, 157, 158 Bruff, J. Goldsborough, 247 Bryant, Edwin, 240 Bryce Canyon, 366 Bubo scandiacus (snowy owl), 173, 175 Bubo virginianus (great horned owl), 172, 173, 175, 179 Buchanan, James, 274 Buddhism, 300 380

Page 380

Buenaventura (mythical river), 89, 93, 97n23 buffalo. See bison buffalo berries, 121 Buffalo Hump, 367, 368 Bullock’s oriole (Icterus bullockii), 172, 173 bullwhackers, 253, 262n1, 320 bunchgrasses, 206 Buntline, Ned, 285 Bureau of Land Management (BLM): and Areas of Critical Environmental Concern, 223; and black trona water, 113; and conservation, 363, 366, 367– 372; and cultural sites, 309n3; and fencing, 366; and horses, 160, 321, 330, 334, 335, 336; and management of Red Desert, 217; and plant species, 228; and roads, 115; and rock art, 250; and uranium prospecting, 349; on vibroseis, 341 Bureau of Reclamation, 113 Burns, R. H., 307, 322 Burris, Dutch Charlie, 357 Burroughs, John Rolfe, 314n4, 322, 326n24, 332, 352n7, 360 burrowing owl (Athene cunicularia), 145, 149, 152, 173, 175, 177 burrows, 151–152, 175, 207 Bush, George W., 336 Bush Rim, 172, 176, 178, 211, 222, 367 bushy-tailed woodrat (Neotoma cinerea), 145, 151 bushy-tailed woodrat (Neotoma cinerea ssp. cinnamomea), 148 Buteo jamaicensis (red-tailed hawk), 173, 175–176

Buteo lagopus (rough-legged hawk), 173, 179 Buteo regalis (ferruginous hawk), 145, 149, 171, 173, 175, 177; nest of, 18 Butler, Anne M., 308 Butler, James, 308 Byrrhidae, 195, 196

C caddis fly, 193 Calcarius mccownii (McCown’s longspur), 144, 146, 149 Calidris minutilla (least sandpiper), 173, 179 Calidris pusilla (semipalmated sandpiper), 173, 179 California gull (Larus californicus), 173, 179 California-Oregon-Mormon Trail: and Cherokee Trail, 254, 255; and conservation, 370; emigrants’ use of, 306; as feature, 305; and freight wagon trains, 253; and gold mining, 343; and Great Divide Basin, 115; inscriptions on Independence Rock, 28; Overland Trail distinguished from, 309n6; and Seminoe Cutoff, 138; and sheep, 317; Sublette Cutoff, 247; and tourists, 217; traffic on, 128, 129 Callaway, S. R., 276, 278 camels, 121, 139, 231 Campbell, Eugene E., 272 Camp Carmichael, 294–295 Camp Douglas, 285–286 Camp Fortunate, 184 Camp Pilot Butte, 267, 275, 276, 279

Index

Camp Stambaugh, 267, 275–276 Canada goose (Branta canadensis), 173, 179 Cane Hill California Emigrating Company, 257 Canella/Canavish (Ute chief ), 287 canine distemper, 160 Canis familiaris (dog), 145, 157 Canis lupus (gray wolf ), 146, 157, 159, 161, 163 canvasback (Aythya valisineria), 173, 179 carbon, 108, 109, 117, 218 –220, 223, 224nn12,15,16, 343 carbonate salts, 111 Carbon County Sheep and Cattle Company, 324 Carbon County Woolgrowers Association, 318 carbon dioxide, 108, 109, 111, 112, 117, 203 Carboniferous period, 343 Cardaria species (hoary cress), 213 Carduelis tristis (American goldfinch), 145, 153, 173, 177 Carex nebrascensis (Nebraska sedge), 221 caribou (Rangifer tarandus), 145 Carley, Maurine, 234 Carmichael, Lewis, 294–295 Carnegie, Andrew, 137 Carnegie Museum, 139 carp (Cyprinus carpio), 145, 157 Carrington (lieutenant), 256 Carroll, Joe, 312 Carson, Kit, 128 Carter, Jimmy, 100n112 Carter, William A., 275 Casa Grande ranch, 319 Casement, Dan, 293–294 Casement, Jack, 293–294 Casper Dunes, 132n17

Proulx_09_pp.377-400:Layout 1

8/4/08

Caspian Sea, 98n45 Cassidy, Butch, 357, 358, 359, 361nn8,15, 361–362n21, 362n24 Cassidy, Mike, 358 Castaño de Sosa, Gaspar, 185n3 Castor canadensis (beaver). See beaver catfish, 109 Catoptrophorus semipalmata (willet), 174, 174 Catostomus commersonii (white sucker), 147, 159 Catostomus latipinnis (flannelmouth sucker), 144, 146, 159 Catostomus platyrhynchus (mountain sucker), 99n83 cattails, 121 cattle (Bos taurus): and disease transmission, 158, 161, 267; feed for, 330; and Irish immigrants, 307; and Little Snake River valley, 311–313; and outlaws, 357–358, 360; and overgrazing, 156, 160, 222, 223, 364; population of, 160, 165n39; and ranching, 317, 318, 319 –321; status in Red Desert, 145 Caucasus Range, 88 Caughey, John W., 239 Cavigelli, J. P., 140 Cedar Canyon rock art site, 246 Cenozoic era, 85, 86, 87, 92 Central Pacific Railroad, 294, 297, 303n2 Centrocercus urophasianus (greater sage-grouse). See greater sagegrouse ceremonial rock art, 245, 246, 247, 248 Cervus elaphus (elk). See elk

3:51 PM

Page 381

Chain Lake horse bands, 330, 333 Chain Lakes Flat, 41 Chain of Lakes, 115 Charadrius montanus (mountain plover), 146, 149, 154, 173, 176, 177, 178 Charadrius vociferus (killdeer), 173, 176 cheatgrass (Bromus tectorum), 157, 159, 161, 212 –213, 219, 222 chemical dating techniques, 245 Chenopodiaceae, 220, 221 Cherokee Advocate, 256 –257 Cherokee Phoenix, 254 Cherokee Rim, Washakie Basin, 40 Cherokees, 255, 257 Cherokee Trail, 254–257, 257, 258, 264n19, 283, 305, 343, 358 Cherry (lieutenant), 288 cherry (Prunus), 273 Cheyenne Club, 307 Cheyenne mobile belt, 90, 97n36 Cheyenne Pass, 343 Cheyennes, 249, 275, 293 Chilton ranches, 319 Chinese Exclusion Act of 1882, 299 Chinese immigrants: and Aspen site excavations, 298 –299; businesses of, 297, 298, 299; as coal miners, 275, 276 –279, 281n38; diet of, 298 –299; and gambling, 299; as gold miners, 297; as railroad workers, 297–298, 299; as waiters, 303n1; weapons of, 299, 303– 304n12 Chinese Massacre, 276 –278, 280nn34–35, 281nn38,45, 286, 299 Chinle Formation, 91, 98n55

Chivington, John, 293 Chivington, L. H. “Doc”, 331, 337n14, 358 –359 chokeberry (Prunus virginiana), 154, 212 Chondestes grammacus (lark sparrow), 171, 173 Chouteau, Pierre, 342 Chrysothamnus (rabbitbrush), 192, 196, 205–206, 212, 220 Chrysothamnus viscidiflorus (Douglas rabbitbrush), 205, 207, 209, 211 Cicadellidae, 196 cinnamon teal (Anas cyanoptera), 171, 173 Circle Bar Formation, 98n65 Circle K ranch, 321–322 Circus cyaneus (northern harrier), 146, 154, 173, 176 Citizens’ Red Desert Protection Alternative, 368, 370 –371 Citizens’ Wildlife and Wildlands Protection Alternative, 370 – 371 Civil War, 274, 275, 285 Clark, D. O., 345 Clark, William, 183–184, 342 Clark’s nutcracker (Nucifraga columbiana), 173, 180 Clay, John, Jr., 326n15, 360, 361n15 clay mud bogs, 115 Clayton, William, 273 Claytor ranch, 319 Cleveland, Grover, 278 cliff swallow (Petrochelidon pyrrhonota), 171, 173, 177 climate: and coal mines, 344; and environmental change, 121, 122, 123, 124, 129 –130, 133n37, 213; and insects, 197; and wildlife, 155

Index

climate regions, 122 Clinton, Bill, 370 Clovis points, 231 Clyman, James, 125, 126, 339 Clymer, John, 352n11 coal-bed methane fields, 162 – 163, 198, 228, 351, 363 Coal Gulch, 93 coal mines: and Chinese workers, 275, 276 –279, 281n38; conditions of, 345–347; development of, 343–344, 363, 367; and environmental change, 128 –129; and Japanese workers, 300, 301, 302 –303, 304n35; and labor strife, 276 –279, 286, 345, 347n6; open-pit and tailings near Jim Bridger Power Plant, 55; and roads, 162; and wildlife, 163 cobble-pebble lag, 21 Coble, John C., 360 Cody, William F., 285 Colaptes auratus (northern flicker), 146, 154, 173, 176, 178 cold war, 348, 350 Coleoptera (beetles), 193 Collins, William O., 268, 284 Collister, Oscar, 283–284 Colorado pikeminnow (Ptychocheilus cius), 145, 159 Colorado Plateau, 92 Colorado River, 92, 93, 94, 99n90, 114, 127 Colorado River Basin Salinity Control Act, 117 Colorado State Recreation Site, 48 Colorow (Comanche chief ), 287, 288 Columba livia (pigeon), 146, 160 Columbian mammoth (Mammuthus columbi), 145, 155 381

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Comanches: and horses, 181– 183, 185n5, 186nn17,19,26,28; and Protohistoric Period, 234, 237n28; and rock art, 246, 249 common grackle (Quiscalus quiscula), 172, 173 common poorwill (Phalaenoptilus nuttallii), 171, 173 common raven (Corvus corax), 145, 148, 171, 173, 179 Como Lake, 261 Comstock, Bill, 285 conglomerate, 91–92, 98nn62,64 conifers, 179, 221 conservation: and Bureau of Land Management, 363, 366, 367– 372; efforts toward, 77, 78, 81, 363–372; and fencing, 161, 366, 367, 369; and hunting, 158, 160, 269n3, 363–364, 366 –367, 368, 370; and national parks and monuments, 217, 364–366; and Native Americans, 368, 370, 371; and open space, 80, 217, 223, 363, 368, 371; of plants, 217, 218, 223, 228; and ranches, 363, 364–365, 366, 367, 369, 371; and trout, 113; and wildlife, 157–159, 160, 161, 163, 166n65, 218, 223, 363–364, 366 –367, 369, 370, 371 continental divide: and AdamsOnís Treaty, 181, 185n1; and environmental change, 126, 133n49; and freight wagon trains, 253; and hydrology, 114–115; and oil and gas production, 372; and Rusling, 259 Continental fault, 90 382

Page 382

Continental Peak, 176, 366 Contopus sordidulus (western wood-peewee), 172, 174, 177 convergent orogeny, 88, 89 Cook, Bob, 77, 78 Cooper’s ranch, 318 Cope, E. D., 87, 137, 138 cordillera of the Americas, 88 – 89, 98n67 corner-notched projectile points, 272 Coronado, Francisco Vásquez de, 181 Corson, Eva McCormick, 323, 327nn29 –30 Corson, John, 323 Cortés, Hernan, 329 Cortez, Sea of, 92 Corvus brachyrhynchos (American crow), 173, 179 Corvus corax (common raven), 145, 148, 171, 173, 179 Corynorhinus townsendii (Townsend’s big-eared bat), 147, 152 Cosgriff, James E., 322 Cosgriff Sheep Company, 322 cottonwood, 126, 153, 154, 196, 221 Courtwright, Frank, 284 Coutant, C. G., 326n14 Cow Creek and Pioneer Sheep ranch, 319, 322 Cow Creek ranch, 319, 322 –324, 327n27 Cragin, Francis Whittemore, 263n5 Craig, James, 290n1 Cranston, Susan Amelia, 241 Crawford, P. V., 241, 256 Creasman, Steve, 232, 233 crested wheatgrass (Agropyron cristatum), 218

Creston Junction, 92, 93 Cretaceous, 90, 138 Crook, George, 286, 288, 291–292n28 Crooks, Ramsey, 342 Crooks Gap Conglomerate, 98n64 Crooks Gap fault, 90 Crooks-Green-Ferris chain, 153 Crooks Mountain, 97n35, 218, 221 Crookston Ranch, 370 Cross, Whitman, 97n20, 100n97 Crotalus viridis (rattlesnake), 147, 149, 152 Crouse, Charley, 362n22 Crows, 233, 234, 248, 249, 340 crustaceans, 174 crust of earth, 87, 88, 89, 90, 96n9, 97n16 cryptobiotic crusts, 227–228 cultural sites, 305, 309n3, 368, 369 culvert installation, 117 Cumming, Alfred, 274 Currant Creek, 125, 127, 254 cutthroat trout (Oncorhynchus clarkii), 144, 145, 159, 160 Cutting, Howard, 129 cyanobacteria, 227 Cyanocitta stelleri (Steller’s jay), 173, 180 Cyclone Rim, 208 Cynomys leucurus (white-tailed prairie dog), 147, 148, 149, 149, 151, 160 Cyprinus carpio (carp), 145, 157

D Dad (ghost town), 31 Daley, William, 360 Daley ranch, 318 dance flies, 194 Danforth, E. H., 287

Index

dark-eyed junco (Junco hyemalis), 173, 177 Darlington, David, 359, 362n26 Dary, David, 157 Deadman Wash Phase, 232, 236n12 deep time, 85, 96n3 deer. See mule deer deer mouse (Peromyscus maniculatus), 145, 148 Dendroica coronata (yellow-rumped warbler), 172, 174, 177 Dendroica petechia (yellow warbler), 174, 177 depocenter, 111, 114 desert side-notched points, 272 DeSmet, Pierre-Jean (priest), 239 Devils Tower, 217 Dines (ghost town), 29 Dinosaur National Monument, 234 dinosaurs, 137, 138 Diplodocus carnegii, 137 Dipodomys ordii (Ord’s kangaroo rat), 146, 150 Dipodomys ordii ssp. priscus (Ord’s kangaroo rat), 148 Diptera (flies), 193 Distichlis stricta (inland saltgrass), 221 Dobie, J. Frank, 329 –330 Dodge, Francis, 289, 291n17 Dodge, Grenville, 291n17 Dodge, Richard I., 286 dog (Canis familiaris), 145, 157 domestic sheep (Ovis aries): areas set aside for, 318, 325nn5–7; and disease transmission, 157, 161; effect on Red Desert, 165n40; forage of, 160, 220; and Little Snake River valley, 311–312; and overgrazing, 80, 156, 160, 318, 325n5; and

Proulx_09_pp.377-400:Layout 1

8/4/08

plant diversity, 222, 225n24; population of, 80, 160, 165n39; and Scots immigrants, 307; status in Red Desert, 145 Domínguez-Escalante expedition, 89, 182, 234, 235n3 Doña Blanca (children’s game), 263n5 Dorn, Robert, 218, 222, 224n13 Dougherty, Tom, 367 Douglas Creek drainage, 123 Douglas fir (Pseudotsuga menziesii), 221 Douglas rabbitbrush (Chrysothamnus viscidiflorus), 205, 207, 209, 211 downcutting, 115–116 Dreyo, Fred, 365 drilling sites, 63, 71 Dripping Rock Springs, 311, 358 – 359, 362n24 Dryinidae, 195 Dunham, Frank, 159, 363–364 Dunis, Doyce B., 239 Durbin, John H., 320 dusky shrew (Sorex monticolus), 144, 145 Dustin, Tom, 366, 367

E Early Archaic period, 232 Earp, Wyatt, 356 earth history, 86 East Africa Rift, 96n8, 97n31 Eastern Shoshone Business Council, 368, 370 Eastern Shoshonis, 234–235, 237n31 East Flaming Gorge, 246 East Pacific rise, 91

3:51 PM

Page 383

Echo Overlook, 7 Eckerle, William, 131nn4,10 ectothermic wildlife, 164n3 Eden, 113 Eden-Farson project, 113 Eden-Farson site, 234 Edmonson (captain), 257, 258 Edwards, E. H., 332 Edwards, Griff, 312, 314n4 Edwards, Jack, 312, 314n4 Eightmile Lake, 44 Einstein, Albert, 87, 350 Eisenhower, Dwight D., 261 Elaeagnus angustifolia (Russian olive), 218 elephants, 139 Eleventh Ohio Volunteer Cavalry, 268, 269n4 elk (Cervus elaphus): hunting of, 267, 287; and migration, 152, 158 –159, 163; and open space, 80; and roads, 161, 162; and rock art, 246, 247; and sagebrush, 153, 209; status in Red Desert, 145; and water, 153; winter game park for, 364 Elk Lake, 109 Elk Mountain, 260, 262, 265n48 Elko series point, 272, 279n4 Elymus cinereus (basin wildrye), 207, 210, 211, 219 Elymus lanceolatus ssp. lanceolatus (thickspike wheatgrass), 206 Elymus smithii (western wheatgrass), 219, 220 emancipation, 291n10 emigrants: and bighorn sheep, 161; characteristics of, 306; diaries of, 306, 309n7; escorts for, 267; and Fort Bridger, 271, 273, 274; and Homestead Act,

254; in Red Desert, 305; and rock art, 247; and Shoshonis, 239 –242, 242 –243n4, 243n16; surge of, 267; trails used by, 254–258, 259, 295, 306, 343; and transcontinental railroad, 259 –260; and U.S. Army, 306, 309n8 Emigrant Trail fault, 90 Empididae, 194 Empidonax minimus (least flycatcher), 173, 177 Encampment River, 311 Endangered Species Act, 159, 163 endemic insects, 197 endemic plants, 218 endemic vertebrates, 148, 149 –150, 151, 152 Endicott, William C., 279n2 endorheic watershed, 115 endothermic wildlife, 144, 152, 164n3 energy extraction work: and biological soil crusts, 228; and changes to Red Desert, 80; and economy, 339, 343–351; and geophysical operations, 340 –341; and Little Snake River valley, 314; and plants, 217, 223; and roads, 162, 305, 341, 363; and wildlife, 163, 343, 363 environmental change: and aridity, 121, 122 –123, 122, 129; and climate, 121, 122, 123, 124, 129 –130, 133n37, 213; effects of, 131n6; and Europeans, 124, 125, 126 –127, 128; and glaciation, 121, 122, 123– 124, 131n4; and grasses, 121, 123, 127, 128, 159; and Little

Index

Ice Age, 122, 123–124, 129, 132nn28,31,32,34; and Native Americans, 121, 124; and soil, 121, 123, 124, 126, 129 –130; and wildlife, 155 Enzi, Mike, 368 Eocene epoch: geology of, 90, 91, 92; and hydrology, 108, 109, 110, 112, 114; paleontology of, 138, 139 Eocene Green River Formation, 109 Eohippus. See Hyracotherium eolian sandstone, 92, 98n66 ephemeral streams, 116 –117, 118n32, 210 Equus caballus (horse). See horses Eremophila alpestris (horned lark), 146, 148, 171, 173 ergs, 90, 98n42 Ericameria (rabbitbrush), 192, 196, 205–206, 212, 220 Ericameria nauseosa (rubber rabbitbrush), 205, 206, 207, 210, 211 erosion: arroyos, 124–125, 127, 128 –129, 156; and biological soil crusts, 227, 228; and environmental change, 121, 124, 129, 130 Erslev, E. A., 97n19 Erwin v. State (1876), 355 Escalante, Silvestre Vélez de, 186n17, 234, 235n3. See also Domínguez-Escalante expedition Espejo, Antonio de, 185n3 Euderma maculatum (spotted bat), 147, 152 Euphagus cyanocephalus (Brewer’s blackbird), 172, 173 Eureka Pool, 324 383

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Europeans: and environmental change, 124, 125, 126 –127, 128; and Protohistoric period, 233, 234–335; on rock art, 245; and wildlife, 156 –157. See also specific nationalities European starling (Sturnus vulgaris), 145, 160 eutrophication, 117 Evans, Lewis, 258 Evans route, 254–256, 257, 258, 264n15 evaporation rate: and Gosiute, 108, 110, 111, 112 –113; and Great Divide Basin, 115; and sagebrush, 203, 209; and wildlife, 153; and winds, 144 evaporative pumping, 111 evaporites, 91, 98n57 Everest, Mount, 97n14 Eversole, Elza, 323, 327n32, 331 Eversole, Roy, 323 Eversole ranch, 319 evolution, 87, 140 Ewers, John C., 182, 184, 248 exotic plants, 161, 212, 218, 271, 273, 274, 307 exotic wildlife, 159 –160 extensional belts, 88

F Falco mexicanus (prairie falcon), 146, 154, 171, 173, 176 Falco sparverius (American kestrel), 172, 173, 176 Faragher, John Mack, 309n7 faults, 90, 94–95 fauna: defined, 164n1; spatial and temporal contexts of, 143. See also birds; fish fauna; wildlife; and specific fauna 384

Page 384

Faure, Gunter, 86 Fautin, R. W., 193 Federal Highway Act, 261 Federal Land Policy and Management Act, 367–368 feldspar, 100n105 fencing: and conservation, 161, 366, 367, 369; junipers as, 359, 362n26 Ferguson, Charles, 78, 338n38 Ferris, Warren Angus, 127, 355 Ferris-Bairoil Road, 30 Ferris Dunes, 95, 100n105, 113, 122, 123 Ferris Mountains, 4; conservation of, 371; geology of, 95; plant diversity of, 218, 221; and sagebrush, 208, 212; and wildlife, 153 ferruginous hawk (Buteo regalis), 145, 149, 171, 173, 175, 177; nest, 18 Fertig, Laura, 78 Fertig, Walter, 78 Figure Four ranch, 318, 320, 321 Finch, Jim, 345 Finerty, John F., 286, 289, 290 Finfrock, John Henry, 284–285, 290nn7– 8 Finlanders, 276, 345, 346 –347 Finley site, 131n2, 231 Firehole Basin No. 11, 234 Firehole Canyon, 117 Firehole Phase, 233, 234 fires: and aspen, 222; and environmental change, 124, 129; and exotic plants, 219, 222; and sagebrush, 205, 206 – 207, 210, 211, 212, 213, 215n39 Fisher, Carl G., 260 Fisher, Fred, 322 fish fauna: of Green River, 92, 93,

94, 99nn82 – 83; peripheral vertebrates, 144; of Platte River, 93, 99n83; poisoning of, 159 –160; in reservoirs, 159; of Snake River, 93, 99n84; and water, 152. See also specific fish Fitzpatrick, Thomas, 239, 339, 340 Flaming Gorge, 92, 117 Flaming Gorge Dam, 247 Flaming Gorge Reservoir, 116, 117, 159 Flaming Gorge rock art sites, 247, 249 –250 flannelmouth sucker (Catostomus latipinnis), 144, 146, 159 “flaring tower,” 64 Flat Top fault, 90 Fletcher, Jack Earl, 256 Fletcher, Patricia K. A., 254–255, 256 flies, 193, 196 Flight of the Wild Stallions (film), 333 flora, vegetation distinguished from, 214n1 fluted points, 231 Fogg, Bobby, 345 fold-thrust belts, 88, 89, 90, 97n17, 98n43 Folsom points, 231 Folsom train robbery, 359 Fontanelle, 372 Food Machinery and Chemical Corporation, 349 –350 forbs: classification of, 214n8; diversity of, 218, 219 –220, 222 – 223; and fire, 206; roots of, 203, 204; and sagebrush, 209, 210 Ford, Henry, 294 foreland uplifts, 88, 90, 91

Index

Fort Bridger: archaeological excavation of, 124, 272, 273–274, 279n4; character of, 267; and environmental change, 121, 124, 131n1, 134n66; and guns, 273–274, 280n11; and Mormon Church, 271, 272, 273, 274; and paleontology, 138; Palmer’s description of, 271–272; restoration of, 291n18; and Shoshonis, 235, 240 –241, 271, 272, 273, 274; and South Pass area, 275; as trading post, 271, 272, 273; and Union Pacific Railroad, 275; and U.S. Army, 272, 273, 274–275 Fort Fred Steele, 267, 275, 286 – 290, 291n18, 291–292n28, 294, 345 Fort Halleck, 267, 283–286, 290n1 Fort Kearney, 286 Fort LaClede, 117, 267, 269n1 Fort Laramie, 268, 269, 286, 291n18 forts, 267–269, 268, 306, 342. See also specific forts Fort Union Formation, 140 fossils: and geologic study, 87; of Green River Formation, 92; and Reed, 137–138. See also paleontology Fourier, Charles, 287 foxtail barley (Hordeum jubatum), 221 Fraeb, Henry, 311, 313 Franzwa, Gregory M., 261, 265n44 Frazer, Robert W., 269n1 freight wagon trains, 253, 262n2, 267, 295 Fremont, John Charles, 89, 128,

Proulx_09_pp.377-400:Layout 1

8/4/08

233, 240, 243n9, 257, 263n14, 264n15 Fremont Glacier, 133n34 Fremont Phase, 233, 235n3, 247, 249 –250 French explorers/settlers, 183, 186nn25–26 Fresnos, 294, 295n1, 313 Freudenthal, Dave, 371 Friends of the Red Desert, 79, 371 Frison, George, 233–234 Fuhr, Paul, 320 Fulica americana (American coot), 173, 179 Fulton, A. D., 325 fulvic acid, 114 fungi, 227 fungus gnats, 195–196 Furniss, Norman F., 272

G gadwall (Anas strepera), 173, 176 gambling, 299 Gannett, J. W., 298 Gannett Peak, 89 Gardner, Dudley, 81, 264n19, 294–295, 348 –349 Gardner saltbush (Atriplex gardneri), 206, 220 gas pipeline: construction in progress near Grenville Dome, 65; forty-two-inch pipeline near Walcott Junction, 67; fortytwo-inch section near Sinclair, 68; intersecting rights-of-way near Bitter Creek, 49; surfacelaid pressurized pipeline, Vermillion Creek drainage, 66; trench right-of-way disturbance east of Adobe Town, 51 Gateway rock art site, 247

3:51 PM

Page 385

General Half-Breed Bill of 1862, 356 geology: and American cordillera, 88 – 89; Cenozoic record, 85, 86; of Gosiute, 90 –91, 92, 94, 95, 98n54; of Green River, 85, 87, 92 –93, 94, 95, 99nn74,90,96; of Green River Formation, 91, 98nn57,58, 100n107; of Leucite Hills, 88, 94–95, 97n20, 100n97; of Neogene, 91–92, 93, 95; orogenesis, 87–90, 96 –97n13; and plant diversity, 218; and sediments, 88, 91, 92, 93–94, 95; and subduction, 87– 89, 90, 91, 96n11, 97n35; table of geologic time, 86; and Wyoming craton, 89 –90 geophysical exploration, 340 –341 Geringer, Jim, 368, 369 German immigrants, 301 Giardia lamblia, 263n13 Gila cypha (humpback chub), 146, 159, 163 Gila elegans (bonytail), 145, 159 Gill, George, 250 glaciation, 121, 122, 123–124, 129, 131n4, 132nn28,31,32,34 Glass, Hugh, 339 Glenn, M., 324 global positioning system (GPS) surveying methods, 341 Gloger’s rule, 150 Goddard, Joe, 348 –349 golden eagle (Aquila chrysaetos), 80, 146, 148, 149, 171, 173, 175, 179 golden-mantled ground squirrel (Spermophilus lateralis), 146 golden-mantled ground squirrel

(Spermophilus lateralis ssp. wortmani), 148 gold mining, 297, 311, 313, 317, 343, 372 gold rush, 271, 317, 343 Gonsales, Jose, grave of, 37 Gooldy, John F., 330 gopher snake (Spermophilus lateralis), 145, 146 Gordon River, 117 Gosiute, Lake: geology of, 90 –91, 92, 94, 95, 98n54; hydrology of, 107–110, 111, 112 –113, 114, 115, 117; maximum extent of, 112 Gosiutes, 272 Gould, E. D., 320 Gould, Jay, 345 Gould, Lewis L., 269n2 Gould, Stephen Jay, 96n3 Gowans, Fred R., 272, 352n7 GPS. See global positioning system grabens, 92, 98n65. See also half grabens graminoids, 203, 204, 214n8, 221, 225n20. See also grasses; sedges granaries, 233, 234 Grand Canyon, 92, 93–94, 99n73 Grand Teton National Park, 158, 365–366 granite, 88, 91, 97n21 Granite Mountain fault, 90 Granite Mountains, 97n40, 112 grasses: and Archaic period, 232; diversity of, 218, 219; and environmental change, 121, 123, 127, 128, 159; and fire, 206; and Paleoindian period, 231; range of, 221, 223; and sagebrush, 202, 205, 206, 209,

Index

219; and water, 153, 203, 212, 219, 224n16 grasshoppers, 20, 190, 193, 194–195, 198n4 gravel deposits, 92, 93, 95, 107, 116 gray horsebrush (Tetradymia canescens), 205 Grayia spinosa (spiny hopsage), 205–206, 207, 220 gray vireo (Vireo vicinior), 144, 146 gray wolf (Canis lupus), 146, 157, 159, 161, 163 greasewood: and insects, 190, 192, 193, 194–195, 196, 197; and shrublands, 221. See also black greasewood Great Basin, 88, 97n18, 149 Great Basin spadefoot (Spea intermontana), 144, 146, 151, 152 Great Basin wild rye, 153, 156 great blue heron (Ardea herodias), 172, 173 Great Divide, 85 Great Divide Basin: and birds, 179; and bison, 366; and conservation, 372; and environmental change, 126; geology of, 92, 93, 95, 98n67; hydrology of, 107, 115–117; paleontology of, 137–138 Great Divide Phase, 232 Great Divide Resource Management Plan, 371 greater sage-grouse (Centrocercus urophasianus): conservation of, 158; hunting of, 267; and insect prey, 155; and open space, 80; predators of, 160; and roads, 161; and sagebrush, 153, 171, 176 –177, 179, 201, 206, 209, 212; status in Red Desert, 146, 149, 172, 173 385

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

greater short-horned lizard (Phrynosoma hernandesi), 148 great horned owl (Bubo virginianus), 172, 173, 175, 179 Great Plains, 89, 149, 160 Great Salt Lake, 108, 110, 122, 123, 254 Great Sand Dunes National Monument, 366 Greek immigrants, 300, 303 Greeley, Horace, 128, 287 Green, Sidney, 320 green algae, 227 Green Mountain, 153, 175, 176, 179, 218, 221 Green River: confluence with Yampa River, 7; and emigrants, 239 –240, 241, 243n9; and environmental change, 125, 126 –127, 128, 129, 134n66, 134–135n72; fish fauna of, 92, 93, 94, 99nn82 – 83; and fish poisoning, 159; geology of, 85, 87, 92 –93, 94, 95, 99nn74,90,96; and hydrology, 107, 116; oil shales of, 110; and rock art, 247, 248, 249, 250; and trappers, 156, 342 Green River Basin: fish fossils of, 92; geology of, 88, 91, 92 –93, 94; and hydrology, 107; King’s map of, 262n5; and paleontology, 138 Green River Formation: fossils of, 92; geology of, 91, 98nn57– 58, 100n107; hydrology of, 109, 112 Green River Resource Area Management Plan, 368 Green River valley: and beaver, 340, 342, 352n6; development 386

Page 386

of, 363; and emigrants, 241; human developments in, 158; and oil and gas fields, 162 green-tailed towhee (Pipilo chlorurus), 171, 173, 178, 212 green-winged teal (Anas crecca), 171, 173, 176 Grieve, Tom, 324 Grieve ranch, 319 grizzly bear (Ursus arctos), 146, 157, 161, 163, 222 Gros Ventre Range, 94, 108 groundwater, 117 growing season, 144 Guangdong Province, 298 gunfighters, 356 –357, 361n9 Gunnison, J. W., 254 Gymnorhinus cyanocephalus (pinyon jay), 146, 154, 173, 180

H haboobs (dust storms), 95, 100n112 Haley, Ora, 313, 319, 320 –321, 326nn21–22 half grabens, 92, 98n65. See also grabens Halictidae, 195 Halleck, Henry W., 290n1 halogeton (Halogeton glomeratus), 157, 161, 213, 219, 222 Hamilton, Bruce, 367 Ham’s Fork, 241 Hansen, Clifford, 364 Hansen, Jim, 330, 333–334 Hansen, Robert, 365 Hansen, W. R., 92 –93 Hardinburgh, John, 313 Hardy, E. N., 334 Harry, Bryan, 366 Hartt, John K., 322, 323 Hastings Cutoff, 240

Hatfield, Nina, 369 Hauser, George, 291n18 Hawthorne, Mont, 345–346 Hay, John, 113, 261, 335 Hay, Leonard, 318, 335, 366 Hayden, Ferdinand V., 87 Hayden expedition, 262n2 Hay Reservoir, 115 Hays, Lorena, 128 Haystacks, 11, 117, 305 Hazen, William B., 264n27 HBC. See Hudson’s Bay Company heavy minerals, 95, 100n107 Hemiptera (bugs), 193, 194 Henry, Andrew, 339, 340 Henry, Guy V., 289, 291n19 Henry’s Fork, 127, 134n54, 139, 341 Herschler, Ed, 367 Hesperostipa comata (needle-andthread grass), 206, 209, 219, 220 Heterodon nasicus (western hognose snake), 144, 147 hibernation, 152, 153, 164n11 hide paintings, 245 High Savery Reservoir, 160 Himalayas, 88, 97n14 Hispanic railroad workers, 303 Historic era, 234, 235n3, 246 hoary cress (Cardaria species), 213 hoary goldenweed (Machaeranthera canescens), 206 Hoback River, geology of, 94 hoboes, 293, 296n2 Hoebel, E. Adamson, 182 Hogan, Everett, 333 Holladay, Ben, 134n66, 283, 290n4, 306, 309n6 Holladay’s Overland Stage line, 254, 283, 306

Index

Hollingsworth, J. J., 283, 285 Holmes, Calvin, 264n19 Holmes, Oliver Wendell, Jr., 355–356, 361n6 Holocene epoch, 122, 123 Homestead Act of 1862, 79, 254 Homestead Act of 1916, 321 Homoptera (aphids, plant hoppers), 193, 195 Honeycomb Buttes, 6, 176, 178, 305, 366, 367, 368 Hooten ranch, 319 Hordeum jubatum (foxtail barley), 221 Horn, Tom, 321, 357, 360 horned lark (Eremophila alpestris), 146, 148, 171, 173 horse-mounted warriors, and rock art, 246, 247, 248 –250 Horse Rescue of America, 166n74 horses (Equus caballus): carcass of, 19; and emigrants, 240; and environmental change, 121, 124, 125, 126, 127; feral horses, 15, 160, 222, 223, 329, 332, 334; and Fort Bridger, 272; free-ranging horses, 159, 329, 330 –331, 334–335, 337n19; horsemeat, 332, 333, 336, 337nn15–17; mustang stock, 329 –330, 335, 336n4; and Native Americans, 155, 181–184, 185nn2,4, 186nn15,19,26,30, 187n40, 235, 237n33, 240, 241, 243nn4,17, 287, 291n21, 329, 330; and outlaws, 358; and paleontology, 139, 231; and plant diversity, 222; and railroad grading, 295; and ranching, 331–332, 333, 337n15; and rock art, 246, 247, 248 –250; run-

Proulx_09_pp.377-400:Layout 1

8/4/08

ning horses, 330 –333, 334, 337n22; and Spanish explorers/settlers, 155, 181–183, 184, 185nn3– 4, 185–186n15, 329, 330; status in Red Desert, 146, 157; traps for, 332 house mouse (Mus musculus), 146, 157 house sparrow (Passer domesticus), 146, 160 house wren (Troglodytes aedon), 172, 173, 177 Hudnell, Joseph, 284 Hudson’s Bay Company (HBC), 127, 342 Hugus, J. W., 291n26, 313, 315n14 Hugus, W. B., 291n26 Hultkrantz, Ake, 234 human occupation: and climate, 121; dating of, 131n2; and vegetation, 213 Humfreville, J. L., 285 humic acid, 114 humpback chub (Gila cypha), 146, 159, 163 Humphrey, Mag, 315n13 hunting: and conservation, 158, 160, 269n3, 363–364, 366 – 367, 368, 370; and forts, 267– 268; and Native Americans, 124, 125, 231, 232, 233, 234, 312, 314n5; and projectile point collection, 79; and stone cairns, 359. See also trappers Huntington, Collis P., 303n2 Hurt, Joel J., 306 Hutton, James, 85, 87, 96n3 hydrology: and Bitter Creek, 116, 117, 119n41; and black trona water, 113–114; current conditions, 114–117; and Gosiute, 107–110, 111, 112 –113, 114, 115,

3:51 PM

Page 387

117; and Green River, 107, 116; and trona, 108, 110 –113, 112, 115 Hymenoptera (ants, bees, and wasps), 193 Hyracotherium, 139

I Ice, Clyde, 333 Ice Point Conglomerate, 98n64 Ichneumonidae, 195 Ickes, Harold L., 364 Icterus bullockii (Bullock’s oriole), 172, 173 Iddings, Joseph P., 100n97 igneous rocks, 85, 94, 100n97 Independence Mountain, 97n37 Independence Road, 255 Independence Rock, 28 Indian Gap Trail, 370 Indian Ocean, 96n8 Indian ricegrass (Achnatherum hymenoides), 206, 209, 219, 220, 241 Indian Road, 306 Indian Wars, 286 Ingalls, Rufus, 264n27 inland saltgrass (Distichlis stricta), 221 insectivorous vertebrates, 154 insects: diversity of, 77, 189 –198; field research on, 191–194, 198n7; as prey, 154–155, 174, 175, 176, 177, 178, 179; and sagebrush, 190, 192, 193, 194–195, 196, 197, 202, 204; and sampling equipment, 191–192, 191, 192, 193, 196, 197; specialization of, 197; and spring, 172; and water, 152 Insley, J. B., 320

integrated gasification combined cycle (IGCC) power plant, 344 Intermittent Stream, 118n32 Interstate 80, 79, 80, 161, 163, 217, 262, 265n48, 305, 308 invertebrates, 154–155, 174 Irish immigrants, 293, 297, 298, 303, 306 –307 irrigation: hydrology of, 116 –117; and plant diversity, 218 –219; reservoir on High Savery, 53 Italian immigrants, 301 Izaak Walton League, 366

J Jack Morrow Canyon, 357 Jack Morrow Creek, 126, 178, 357 Jack Morrow Hills: and birds, 175; conservation of, 363, 368 –372; as feature, 305; namesake of, 357; and sagebrush, 211, 212; and wildlife, 153, 159 Jack Morrow Hills Coordinated Activity Plan, 368 Jack Morrow Hills Study Area, 363, 368, 369, 370, 371 Jackpot uranium mine, 351 jackrabbits, 231, 232. See also white-tailed jackrabbit Jackson, David, 339 Jackson, Donald, 264n14 Jackson, William Henry, 253, 254, 262n2 Jackson Hole National Monument, 364–365 jadeite, 90, 97n35 Japanese immigrants: businesses of, 302; as coal miners, 300, 301, 302 –303, 304n35; as railroad workers, 300 –302; and World War II, 301–302

Index

Jarvie ranch, 319 Jawbone ranch, 319, 322, 324–325 Jefferson, Thomas, 339, 342 Jeffrey, C. W., 350 Jenks, G. D., 285 Jennings, Bob, 285 Jerusalem crickets, 192, 197 jet stream, and environmental change, 121, 123 Jim Bridger Power Plant, 344, 367; cooling ponds at, 56; smoke plumes of, 57 Joe Hay Rim, 222 John, James, 128, 239 –240 Johnson, Coffee, 346 Johnson, David E., 231 Johnson, Hervey, 306, 309n8 Johnson, Lady Bird, 265n48 Johnson County War, 360 Johnston, Albert Sidney, 274, 280n17 Johnston, Clarance T., 313, 314n8 Johnston, James, 291n18 Johnston, Velma, 334 Joliet, Irene, 350 Jonah Field, 372 Jones, George, 191 Jones, William Claude, 256 Jones Route, 254, 256 –257, 264n24 JO ranch, 319, 322, 324 Jorgensen, C. D., 190 Joy, Helen, 261 Joy, Henry B., 260, 261, 265n46 Junco hyemalis (dark-eyed junco), 173, 177 Juncus balticus (Baltic rush), 221 junipers: and birds, 179; as fencing, 359, 362n26; and insects, 194; and temperatures, 221 Juniperus osteosperma (Utah juniper), 221 387

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Juniperus scopulorum (Rocky Mountain juniper), 221 Junk Hill, 95 J. W. Hugus and Company, 313

K Kaibab Uplift, 92 Karel ranch, 318 Kawaguchi, Tom, 302 Kelley, Helen, 315n13 Kemble, Fanny, 290n7 Kentucky bluegrass (Poa pratensis), 218 kerogen, 109, 110 Keyser, Jim, 245, 247, 249 killdeer (Charadrius vociferus), 173, 176 Killpecker Creek, 85, 99n87, 117, 129 Killpecker Dunes 16, 17; and birds, 174, 175; and environmental change, 126; geology of, 94, 95, 100n105; and insects, 154, 192, 195–196, 197; plant diversity of, 220 –221; and rock art, 248; and sagebrush, 209 –210, 211; and Steamboat Mountain, 85; and wind, 144 Kincaid, Gordon G., 337n15 King, Clarence, 87, 262n5 Kinnaman, Daniel L., 264n14 Kinney, Charles, 258, 264n25 Kinney, Tim, 306, 321–322 Kinney ranch, 319 Kinney Rim, 305 kit fox (Vulpes macrotis), 144–145, 146 Kittson, William, 126 Knight, Dennis H., 202, 327n26 Knight, Samuel, 87 388

Page 388

Knight, Wilbur, 137, 139 Knights of Labor, 276 Koehler, Bart, 367, 369 –370, 372 Kootenais, 249 Krakatau, 133n34 Krascheninnikovia lanata (winterfat), 220, 223, 224n17 Krmpotich site, 131n2 Kurosawan fog banks, 85

L L7 ranch, 313, 318, 319 –320 LaBarge–Big Piney rock art site, 247 LaBarge Bluffs rock art site, 247, 248, 249 lacustrine sediments, 108 lacustrine species, 159 lacustrine strata, 91 lake trout (Salvelinus namaycush), 146, 159 Lakota Sioux, 306 Lamar, Howard R., 186 –187n40 Lamb, Bruce, 362n26 Lamoreaux, Roy, 333 Lamoureaux, Oliver, 275–276 Lancelot Eden points, 231 Langum, David J., 361n5 Lanius ludovicianus (loggerhead shrike), 171, 173 Lapinskas, Mark, 314 Laramide, 90, 108 lark sparrow (Chondestes grammacus), 171, 173 Larus californicus (California gull), 173, 179 Larus delawarensis (ring-billed gull), 173, 179 Lasionycteris noctivagans (silverhaired bat), 147, 152

Late Archaic period, 232, 236n12 Late Prehistoric period, 232, 233– 234, 235n3, 249 lava flows, 94–95 Lavender, David, 343 La Vérendrye, Pierre Gaultier de Varennes, Sieur de, 182, 183, 186n26 Lawrence, Taylor, 367 Lay, William Ellsworth “Elza,” 348, 358, 359 –360 Lay, Marvell, 359 Lay, Maude, 359 lazuli bunting (Passerina amoena), 173, 177 leafhoppers, 169 League of Associated Sportsmen (LAS), 363–364 least chipmunk (Neotamias minimus), 146 least chipmunk (Neotamias minimus ssp. minimus), 148 least flycatcher (Empidonax minimus), 173, 177 least sandpiper (Calidris minutilla), 173, 179 Leavenworth, Henry, 340 Leavenworth Cattle Company, 313, 318, 319 ledger art, 245, 249 Leggett, E. W., 362n24 Leidy, Joseph, 138 Lemmiscus curtatus (sagebrush vole), 147, 152, 153 lemon scurfpea (Psoralidium lanceolatum), 210, 220 lentic water bodies, 117 Leopold, Luna, 116 Lepidoptera (butterflies and moths), 193 Lepus americanus (snowshoe hare), 147, 153

Index

Lepus townsendii (white-tailed jackrabbit), 147, 148, 153 Leshy, John, 369 Leslie, Frank, 261, 303n1 Leslie, Mrs. Frank, 299 lesser yellowlegs (Tringa flavipes), 173, 179 Leucite Hills, geology of, 88, 94–95, 97n20, 100n97 Lewis, Bill, 319, 331 Lewis, Ed, 285 Lewis, Meriwether, 183–184, 342 Lewis, M. G., 275 Liberty, Margot, 184 lichens, 227 light traps, 191–192, 192, 197 Lillegraven, Jason, 96n2, 138, 139 –140 limber pine (Pinus flexilis), 221 limestone, 90, 91, 208 limnology, 109 Limosa fedoa (marbled godwit), 173, 174, 179 Lincoln Highway, 38, 260 –261, 262, 265nn37,44,46 Lincoln Highway Association, 260, 261 Lisa, Manuel, 339 lithic manufacturing, 79, 231 lithification, 112 Little Firehole Canyon, 12 Little Ice Age, 122, 123–124, 129, 132nn28,31,32,34 Little Pine–Cold Spring mountain country, 153 little sagebrush (Artemisia arbuscula ssp. arbuscula), 202 Little Sandy Creek, 239 Little Sandy River, 95, 115, 128 Little Snake River, 93, 116, 255, 258 Little Snake River valley: and

Proulx_09_pp.377-400:Layout 1

8/4/08

cattle-sheep ranch conflicts, 311–312; and energy extraction work, 314; investment scams, 313, 314n8; roads of, 313–314, 314–315n10, 315n11; settlement of, 311–312; and trappers, 311, 314n1; and White River Utes, 312 livestock: and aspen, 222; and diseases, 157, 160 –161; and environmental change, 128; poisoning of, 219; and ranching business, 307–308; and sagebrush, 213; and water, 153, 160; and wildlife, 156 –157, 158, 160 –161, 162. See also specific animals Lockwood, Jeff, 77–78, 79 Lockwood, Ron, 322 lodgepole pine (Pinus contorta), 221 Lodore, Canyon of, 92 Logan, Alec, 331 Logan, William B., Jr., 318, 326nn10 –11, 327n30, 330, 332 Logan’s ranch, 319 loggerhead shrike (Lanius ludovicianus), 171, 173 Long, Carol, 371 Long, Max, 366 –367 Longabaugh, Harry, 357 long-billed curlew (Numenius americanus), 173, 179 long-eared owl (Asio otus), 173, 175 Long Valley caldera, 94 Loosle, Byron, 76 Lost Creek horses, 330 Lost Soldier Creek, 125 Louisiana Purchase, 339 Love, Charlie, 85 Love, David, 99n76, 156, 347, 366, 367 Love, Tex, 80, 331

3:51 PM

Page 389

Lowie, Robert H., 237n31 Loxia curvirostra (red crossbill), 147, 153 Luman’s ranch, 319 lungworm, 161 Lupinus pusillus (rusty lupine), 220 Lygaeidae (seed bugs), 193, 194 Lynx rufus (bobcat), 145

M MacDonald, Thomas Harris, 261 MacGillivray’s warbler (Oporornis tolmiei), 173, 177 Machaeranthera canescens (hoary goldenweed), 206 Mackenzie, Alexander, 97n22 Mackenzie Mountains, 88, 97n22 MacPhee, Danny, 97n31 magmatic arcs, 88, 89, 90 maize, 233, 234 Malaise traps, 191, 191, 192, 197 mallard (Anas platyrhynchos), 171, 173, 176, 179 mammals: and Archaic period, 232; and migration, 152; and Paleoindian period, 231; peripheral vertebrates, 144; Pleistocene, 124, 155; widespread vertebrates, 148. See also specific animals mammoth, 121, 231 Mammuthus columbi (Columbian mammoth), 145, 155 Man Butte, 305 Mandans, 182, 183, 186n32, 352n4 mantle (layer of earth), 94, 100n98 Mantodea (mantids), 193 maps: and Clarence King, 262n5; of ocean floor, 87, 96n8; and Stansbury, 254, 262 –263n5;

and Works Progress Administration, 352n8 marbled godwit (Limosa fedoa), 173, 174, 179 Marcy, Randolph B., 129, 274 Marsh, Charles S., 291n21 Marsh, O. C., 87, 137, 138 Marsh-Cope feud, 87 Martel, Wes, 368 Martin, Alejo, 182 Maupin, Louis, 348 Maxon ranch, 318 Mayes, Samuel Houston, 257 McCarty, Bill, 358, 361nn17,19 McCarty, Tom, 358, 361n19 McClure, Charles, 261 McCown’s longspur (Calcarius mccownii), 144, 146, 149 McDermot, John, 113 McDermott, John, 275 McGath ranch, 318 McGee, Gale, 366 McKay Lake, 179 McKinstry, Byron, 241 McLynn, Frank, 265n35 McNary, Tim, 190 meandering streams, 210 –211 Medicago sativa (alfalfa), 218 Medicine Bow, 260, 262 Medicine Bow Crossing, 283 Meeker, Nathan, 286 –288, 289, 291nn22,24, 291–292n28 Meeker Massacre, 286, 288 – 290, 289, 291n20, 292nn29, 31, 312, 324 Melanoplus bowditchi (sagebrush grasshopper), 194 Meldrum, Bob, 360 Mephitis mephitis (striped skunk), 147, 160 Merritt, Wesley, 289, 292n31 Mesa Verde cliff dwellings, 262n2

Index

Mesoproterozoic era, geology of, 91, 97n38 Mesozoic era, geology of, 90, 95 Metcalf, H. H., 359 Metcalf, Michael, 232, 233 methane, 108, 117, 162 –163, 198, 228, 344, 351 Mexican-American War, 267, 271, 317, 341–342 Mexican Revolution of 1821, 185n1 Meyer, L. R., 347n6 micaceous rocks, 97n17 Michael, Howard, 283 microhabitats, 79 Middle ranch, 318 Middlesex Land and Cattle Company, 322 midocean ridge, 87, 89, 96n10 Midway Coal Mining Company, 324 migration: of birds, 172, 174, 175, 177, 178, 179; of wildlife, 152, 158 –159, 161, 363, 366 Milankovich’s cycles, 121 Milk Creek, battle of, 286, 288 – 290, 289, 291n20, 292nn29,31, 312, 324 milkweed bug, 194 Miller, James, 133n37, 135n74 Miller, Leslie, 159, 364 Miller, Mark E., 291n20 Mimbres Basin, 98n67 mineral salts, 117 Miners Delight, 275 mining, 54; biological soil crusts, 228; and conservation, 368, 369; gold mining, 297, 311, 313, 317, 343, 372; and plants, 223; for trona, 349 –350, 363, 367; uranium mining, 62, 117, 348 –349, 350; and wildlife, 143. See also coal mines 389

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Miocene epoch: geology of, 92, 93, 94; paleontology of, 138, 139 Miocene-Pliocene transition, 92, 94 Mionczynski, John, 367 Mitchell, James, 257, 258, 264n23 Mitchell-Quesenbury group, 257 Mizner, Henry R., 286 mobile belts, 89 Montgomery, Milford, 330 Moody, Ralph, 264–265n32 Moonstone Formation, 98n65 Morgan, Dale L., 133n49, 264n14 Morgan, Helen, 337n12 Morgan, Mariah, 315n13 Morlete, Juan, 185n3 Mormon Church, 271, 272, 273 Mormon Trail fault, 90 Mormon War, 255, 272, 274 Morris, Isaac, 296 Morrow, H. A., 275 Morrow, Jack, 357 mosquitoes, 127, 134n54 mosses, 227 Motel Yes No, 36 mountain ash (Sorbus), 273 mountain big sagebrush (Artemisia tridentata ssp. vaseyana), 202, 203, 210, 211–212, 213, 219 mountain bluebird (Sialia currucoides), 173, 178 mountain fever, 255, 263n13 Mountain Fuel Supply Company, 349 mountain lion (Puma concolor), 146, 161 mountain plover (Charadrius montanus), 146, 149, 154, 173, 176, 177, 178 mountain sheep. See bighorn sheep 390

Page 390

mountain silver sagebrush (Artemisia cana ssp. viscidula), 202, 210, 211 mountain sucker (Catostomus platyrhynchus), 99n83 Moxa Arch, 372 Muddy Creek, 4, 93, 116, 125, 133n49, 254 Muddy Gap, 125 Muddy Trail, 313, 315n11 mudflats, 110, 111, 115 mudstone, 91, 92, 217 mule deer (Odocoileus hemionus): hunting of, 233, 267; migration of, 152; and oil and gas production, 161; and roads, 162; and sagebrush, 154, 201, 206, 208, 209, 211, 212, 213; status in Red Desert, 146; thriving populations of, 158; and water, 153 Murcray, D., 362n26 Murie, Mardie, 365 Murie, Martin, 365 Murie, Olaus, 365 muskrat (Ondatra zibethicus), 146, 152 Mus musculus (house mouse), 146, 157 Mustela nigripes (black-footed ferret), 145, 149, 160, 163 Mycetophilidae, 195–196 mycorrhizal associations, 204, 205 Myers, W. A., 348 Myllimaki, Jacob, 345 Myotis ciliolabrum (small-footed myotis), 147, 152

N Naftz, David, 123, 132nn31,34 Nagy, Julius, 202

Names Hill rock art site, 247 Napier Complex, 97n32 National Elk Refuge, 159 National Environmental Policy Act, 369 National Outdoor Leadership School (NOLS), 371 National Park Service, 291n18, 366, 367 National Wildlife Federation, 367, 369 Native Americans: artifacts of, 305; and Jim Bridger, 271; and conservation, 368, 370, 371; and emigrants, 239 –242, 242 –243n4, 243n16, 255, 259, 263–264n14, 306; and environmental change, 121, 124; and Fort Bridger, 235, 240 – 241, 271, 272, 273, 274; and General Half-Breed Bill, 356; and horses, 155, 181–184, 185nn2,4, 186nn15,19,26,30, 187n40, 235, 237n33, 240, 241, 243nn4,17, 287, 291n21, 329, 330; and Indian Wars, 286; and miners, 275–276; and Mormons, 272; and Paleoindian period, 231; removal of, 342; and reservations, 267; and rock art, 233, 245–250; and Spanish missions, 317; and telegraph, 283–284; trails of, 343; and trappers, 125, 340, 342 –343, 344, 352n14, 355; troops’ battles with, 268 –269. See also specific tribes Navajos, 249 Nebraska sedge (Carex nebrascensis), 221 needle-and-thread grass (Hesper-

Index

ostipa comata), 206, 209, 219, 220 Neiman, Charley, 319 –320 Nelson, Aven, 77, 208, 217, 218, 222, 225n24 Neogene period, geology of, 91– 92, 93, 95 Neoproterozoic era, geology of, 91, 97n38 Neotamias minimus (least chipmunk), 146 Neotamias minimus ssp. minimus (least chipmunk), 148 Neotoma cinerea (bushy-tailed woodrat), 145, 151 Neotoma cinerea ssp. cinnamomea (bushy-tailed woodrat), 148 Neuroptera (lacewings), 193 newspaper correspondents, 286, 287, 289 –290, 294–295, 346 Nez Percé, 249 Nicaagat (Ute Jack), 287–288, 291nn19,25 Nichols, Willie, 360 Niland, John, 325 Niland, William, 324 Nishi, George, 300 nitrate, 109 Noble, Gale, 366 Northern Arapaho Business Council, 370 northern flicker (Colaptes auratus), 146, 154, 173, 176, 178 northern goshawk (Accipiter gentilis), 146, 153 northern harrier (Circus cyaneus), 146, 154, 173, 176 northern leopard frog (Rana pipiens), 148 northern pintail (Anas acuta), 171, 173, 176

Proulx_09_pp.377-400:Layout 1

8/4/08

northern pocket gopher (Thomomys talpoides ssp. ocuis), 148 northern pocket gopher (Thomomys talpoides), 146 northern shoveler (Anas clypeata), 173, 176, 179 North Platte River, 115, 133 North Table Mountain, 94 North West Company, 126, 342 Northwest Passage, 97n22 Norton, Gale, 369 Norway rat (Rattus norvegicus), 146, 157 Noyes, Stanley, 185n5 Nucifraga columbiana (Clark’s nutcracker), 173, 180 Numenius americanus (long-billed curlew), 173, 179

O oatmeal trails, 192, 197 ocean floor, mapping of, 87, 96n8 oceanic allegiances, 92 –93, 95 oceanic crust, 88, 90, 97n16 Odocoileus hemionus (mule deer). See mule deer Odonata (dragonflies and damselflies), 193 O’Ferrell, John, 283, 290n2 Ogden, Peter Skene, 89 oil and gas production: and Bureau of Land Management, 368, 369, 370; development of, 344, 347–349, 363, 372; and geophysical operations, 340 –341; and hydrology, 117; and plants, 222; and roads, 162; and wildlife, 143, 158, 161 oil shale, 109, 110, 114, 295 oil shale laminite, 91, 98n54 Okano, George, 302 –303

3:51 PM

Page 391

Old Spanish Trail, 239 Oligocene epoch: geology of, 91; paleontology of, 138, 139 olive-backed pocket mouse (Perognathus fasciatus), 146, 150 – 151 olive-backed pocket mouse (Perognathus fasciatus ssp. callistus), 148 olive-backed pocket mouse (Perognathus fasciatus ssp. litus), 148 Olsen, Pete, 337n12 Omahaws, 241 Oñate Salazar, Juan de, 181, 182, 185n3 Oncorhynchus clarkii (cutthroat trout), 144, 145, 159, 160 Oncorhynchus mykiss (rainbow trout), 147, 159 Ondatra zibethicus (muskrat), 146, 152 Onís Gonzalez Vara, Luis de, 185n1 Opal Phase, 232 open space: and birds, 177; and conservation, 80, 217, 223, 363, 368, 371; and wildlife, 80 ophiolites, 96n10 Oporornis tolmiei (MacGillivray’s warbler), 173, 177 Opuntia polyacantha (prickly-pear cactus), 206 Orabona, Andrea, 78, 81 Ora Haley Company, 313, 319 orange-crowned warbler (Vermivora celata), 146, 153, 173, 177 Ord’s kangaroo rat (Dipodomys ordii), 146, 150 Ord’s kangaroo rat (Dipodomys ordii ssp. priscus), 148 Ordway, Edward, 286, 294, 295n3 Oregon boundary dispute, 267, 341–342

Oregon Buttes, 175, 179, 305, 366, 368, 370, 372 Oreoscoptes montanus (sage thrasher), 147, 153, 171, 173, 177 ornate tree lizard (Urosaurus ornatus), 146 orogenesis, 87–90, 96 –97n13 Orthoptera (grasshoppers and crickets), 193 Osborne, John, 357 Ostermann, Harry, 261 Otoes, 241 Ouachito orogeny, 90 Ouray (Ute chief ), 290, 292n31 outlaws: and cattle rustling, 357– 358, 360; depictions of, 361n8; and gunfighters, 356 –357; hideouts of, 79, 311, 358 –359, 362nn22 –24; and ranchers, 312, 358; and Wild Bunch, 358, 359, 361n19 overburden pressure, 114 overgrazing: and cattle, 222, 223, 364; and environmental change, 124, 128, 129; and grasses, 219; management of, 80, 160; and plant diversity, 222 –223, 225n24; and ranching business, 311; and sagebrush, 212, 215n38; and sheep, 318, 325n5, 364; and water, 117; and wildlife, 156, 157, 160 Overholtz, Miles, 312 Overland Mail, 274, 285 Overland Stage, 258 –259 Overland Trail, 128, 134n66, 274, 275, 283, 286, 295, 309n6, 313, 343 Ovis aries (domestic sheep). See domestic sheep Ovis canadensis (bighorn sheep),

Index

145, 153–154, 157, 158, 160 –161, 222, 367 Owens, Dick, 255, 264n25

P Pacific Creek–Little Sandy River area, 212 Pacific Rim, 88 Pacific Springs, 128 Paiutes, 272 Paleo/Archaic transition, 231 Paleocene epoch, 138 –139, 140 Paleogene period: geology of, 90, 92, 95; sediments of, 107, 108 Paleoindian period, 231, 235n3 paleontology: and Reed, 137–138; and Union Pacific Railroad, 138 Paleoproterozoic era, 89 Paleozoic era, 90 Palmer, Joel, 241, 271–272 parasites, 195, 197 Pardington, A. R., 260 Parker, Robert LeRoy (Butch Cassidy), 357, 358, 359, 361nn8,15, 361–362n21, 362n24 Parrott, Big Nose George, 357 Parrott gang, 357 Passerculus sandwichensis (savannah sparrow), 173, 177 Passer domesticus (house sparrow), 146, 160 Passerina amoena (lazuli bunting), 173, 177 pasteurella, 161 Pastor, Jana, 231 Pathfinder Reservoir, 159 Patton, Bill, 321 Pawnees, 182, 183, 186n25, 241 Payne (captain), 288 peat bog, 109 –110 391

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

pediments, 108 Pegler, Westbrook, 364 Pelecanus erythrorhynchos (American white pelican), 172, 173, 179 Pennock, Taylor, 291n26 Penrose, Charles B., 361n9 People for Wyoming, 368 perennial streams, 116 Perilampidae, 195 peripheral plants, 218 peripheral vertebrates, 144, 144, 149 periphyton, 109 Perkins, Charley, 360 Permanent Indian Frontier, 342 Permanent Wyoming Mineral Trust Fund, 343 Perognathus fasciatus (olive-backed pocket mouse), 146, 150 –151 Perognathus fasciatus ssp. callistus (olive-backed pocket mouse), 148 Perognathus fasciatus ssp. litus (olive-backed pocket mouse), 148 Peromyscus maniculatus (deer mouse), 145, 148 Perryman, Berry, 205 Peterson, Jepp, 330 Peterson, Kenneth, 123 Peterson, Olaf, 137, 139 Petrochelidon pyrrhonota (cliff swallow), 171, 173, 177 petroglyphs, 25, 245–246, 247, 248, 249, 305, 370 Peugh, Jack, 367 Péwé, Troy, 95 Phalaenoptilus nuttallii (common poorwill), 171, 173 Phalaropus tricolor (Wilson’s phalarope), 174, 174 392

Page 392

Phasianus colchicus (ring-necked pheasant), 147, 160 phosphates, 109, 350 phosphorus, 117 Phrynosoma hernandesi (greater short-horned lizard), 148 phytoplankton, 109 Pica hudsonia (black-billed magpie), 173, 179 Pick, Vernon, 350 Picket Creek, 137 Picket Lake, 137–138, 139, 179 pictographs, 245, 246, 249 Pierson, Al, 369 pigeon (Columba livia), 146, 160 pill beetles, 196 Pilot Butte, geology of, 94 Pine Butte Ranch, 319, 327n30 Pine Canyon, 179 Pine Canyon rock art site, 246, 249 Pine Mountain, 112, 263n5 Pine Spring Phase, 232, 236n12 Pine Springs site, 231 pine trees, 121, 176, 221 Pinkerton, 356 Pinus contorta (lodgepole pine), 221 Pinus flexilis (limber pine), 221 pinyon jay (Gymnorhinus cyanocephalus), 146, 154, 173, 180 pipeline corridors, 210, 213, 222, 228, 305, 341, 363 Pipilo chlorurus (green-tailed towhee), 171, 173, 178, 212 Pipunculidae, 196 Piranga ludoviciana (western tanager), 172, 174, 177 Pirsson, Louis V., 100n97 pitfall traps, 191, 192, 196 pit houses, 232 Pitkin, Frederick W., 290 Plains of San Agustín, 98n67

plains silver sagebrush (Artemisia cana ssp. cana), 211 plants: adaptations of, 219 –222; and biological soil crusts, 227; and burrowing animals, 152; changes in, 222 –223; conservation of, 217, 218, 223, 228; diversity of, 194, 217, 218 – 220, 221, 222 –223, 224n3, 225n24; riparian vegetation, 153, 154, 156, 157, 164n13, 177, 194, 219, 221. See also specific plants plate tectonics, 88, 89, 90, 91, 92, 94, 96n11 Platte River: and environmental change, 126, 127; fish fauna of, 93, 99n83; geology of, 87, 92 –93 playa, 46 playa lake model, 108 Pleasant Valley Coal Company payroll heist, 359 Plegadis chihi (white-faced ibis), 174, 179 Pleistocene epoch, geology of, 92 –93 Pliocene epoch: geology of, 99nn90,96; paleontology of, 138 Pliocene-Pleistocene transition, 94 Plumb, P. B., 269n1 Plummer, Mitchell, 123 Poa (bluegrass), 219 Poa pratensis (Kentucky bluegrass), 218 Poa secunda (Sandberg bluegrass), 206 pocket habitats, 77–78 Poecile atricapillus (black-capped chickadee), 173, 177, 179 Poetschat, George, 247, 249

Index

Polioptila caerulea (blue-gray gnatcatcher), 173, 178 ponds: and beaver, 125, 126; and birds, 152 –153, 174; and plant diversity, 220, 221; and wildlife, 152 –153 Pooecetes gramineus (vesper sparrow), 171, 174, 177 Populus tremuloides (aspen). See aspen Porter, Fitz-John, 274, 280n17 Pothook ranch, 313, 318, 320 Potter, Carroll H., 285 pottery, 233, 234 Powder Mountain rock art site, 246 Powder Rim, 90 Powder River Basin, 162, 363 Powder Springs, 25, 311, 358 –359, 362n24 Powell, John Wesley, 87, 117, 341 Pownall, Herbert, 366 prairie falcon (Falco mexicanus), 146, 154, 171, 173, 176 precipitation: and Archaic period, 232; and environmental change, 123, 129, 134– 135n72; and hydrology, 110, 115, 116; and plants, 220, 221; and soil water, 207; and wildlife, 143, 144, 145 Prehistoric population, 233, 234 Prehistoric Rim, 91, 305 Preuss, Charles, 254, 263–264n14, 265n35 Price, Jake, 330 –331, 335, 337n22 prickly-pear cactus (Opuntia polyacantha), 206 Pritchard, James A., 241 Procyon lotor (raccoon), 147, 160 projectile points, 79, 231, 233, 234, 272

Proulx_09_pp.377-400:Layout 1

8/4/08

Promontory Point, 286, 294, 296 pronghorn (Antilocapra americana): and Archaic period, 232; and diseases, 157; and environmental change, 121, 128; and fencing, 367; hunting of, 233, 234, 267; and Late Prehistoric period, 233, 234; and migration, 152; predecessors of, 139; Red Desert as range for, 366; and roads, 161, 162, 163; and sagebrush, 153, 154, 206, 208, 209, 212; status in Red Desert, 147, 157, 158, 159; winter game park for, 364 Proterozoic orogenies, 89 –90 Protohistoric period, 233, 234– 235, 249 Prunus (cherry), 273 Prunus virginiana (chokeberry), 154, 212 Pryor Mountain Free-Roaming Horses and Burros Act, 334 Pseudacris triseriata (western chorus frog), 147, 148 Pseudotsuga menziesii (Douglas fir), 221 Psoralidium lanceolatum (lemon scurfpea), 210, 220 Ptychocheilus cius (Colorado pikeminnow), 145, 159 Pulley, Tom, 331 pulsed recruitment, 205 Puma concolor (mountain lion), 146, 161 Purshia tridentata (antelope bitterbrush), 211, 212 Pyeatt (diarist), 255, 256 pygmy rabbit (Brachylagus idahoensis), 144, 147, 149, 152, 153, 206, 208

3:51 PM

Page 393

pygmy sagebrush (Artemisia pygmaea), 202 Pyrenees, 88

Q quartzite, 90, 91, 97nn21,38 Quealy, John, 345 Quealy, Michael, 345 Quesenbury, William, 257–258, 264n23 Quien Hornet Mountain, 254, 263n5 Quinkent, Douglass (Ute chief ), 287, 288, 289, 290, 291n24 Quiscalus quiscula (common grackle), 172, 173

R rabbitbrush (Ericameria or Chrysothamnus species), 192, 196, 205–206, 212, 220 raccoon (Procyon lotor), 147, 160 radiocarbon assays, in Douglas Creek drainage, 123, 132n25 radiocarbon dates, 232 radium, 350 railroad: and black workers, 293; Civil War veterans as laborers for, 293, 297; and grading, 294–295; and hoboes, 293, 296n2; and plants of Red Desert, 217; and rock art, 247; transcontinental railroad, 115, 247, 256, 259 –260, 265n37, 297, 303n2, 356; and troop movement, 267, 286; and wildlife, 162. See also Union Pacific Railroad rainbow trout (Oncorhynchus mykiss), 147, 159

rain shadow, 121, 129, 143 Rana pipiens (northern leopard frog), 148 ranches and ranching: and cattle, 317, 318, 319 –321; and cattle rustling, 357–358, 360; cattlesheep ranch conflicts, 307, 311–312, 322; Circle K ranch, 321–322; and conservation, 363, 364–365, 366, 367, 369, 371; Cow Creek ranch, 319, 322 –324, 327n27; establishment of, 79, 317; and fencing, 366; and Great Depression, 307–308; and horses, 331– 332, 333, 337n15; Jawbone ranch, 319, 322, 324–325; JO ranch, 319, 322, 324; Leavenworth Cattle Company, 313, 318, 319; L7 ranch, 313, 318, 319 –320; and Little Snake River valley, 307, 311–313, 319, 321; and roads, 307; and sheep, 317–318, 321, 322 –325; and Spanish missions, 317; stories of, 319, 326nn10,11; Two Bar ranch, 318, 320 –321, 326n24, 357, 360, 361n15 Randall, Dick, 367 Rangifer tarandus (caribou), 145 Rankin, James, 292n29, 324 Rankin, Joe, 288, 289, 292n29, 324 Rankin, M. Wilson, 292n29 Rankin, Robert, 292n29 Rankin, William, 320 Rash, Madison, 321 Ratliff, Harry, 321 rattlesnake (Crotalus viridis), 147, 149, 152 Rattlesnake Hills, 139 Rattus norvegicus (Norway rat), 146, 157

Index

Rawlins, 34 Rawlins-Bairoil Road, 4 Rawlins Uplift, and sagebrush, 208, 212 razorback sucker (Xyrauchen texanus), 147, 159 Reader, Al, 321 Reader family, 313 recreation, 217, 223, 363–364 Recurvirostra americana (American avocet), 173, 174, 179 red beds, 110 red-breasted nuthatch (Sitta canadensis), 173, 177 red clover (Trifolium pratense), 218 Red Creek Badlands, 368 Red Creek ranch, 319 red crossbill (Loxia curvirostra), 147, 153 Red Desert: cautions concerning, 78 –79; and global environment crisis, 80; myths concerning, 77; settlements of, 307–308. See also conservation Red Desert Ranch, 318 red fox (Vulpes vulpes), 147, 160 Red Lake, 115, 368 red squirrel (Tamiasciurus hudsonicus), 147, 153 red-tailed hawk (Buteo jamaicensis), 173, 175–176 red-winged blackbird (Agelaius phoeniceus), 172, 173 Reed, Bernard, 128 Reed, Bill, 137–138, 139, 140 Reed, Chuck, 330 rendezvous, 127, 134n54, 184, 248, 341, 342, 344, 352n7 Repcheck, Jack, 96n3 reptiles: peripheral vertebrates, 144; widespread vertebrates, 148 restoration ecology, 197 393

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Reynolds, A. W., 157 Richards, W. A., 360 Richards Gap, 247 Rife, E. H., 362n25 ring-billed gull (Larus delawarensis), 173, 179 ring-necked pheasant (Phasianus colchicus), 147, 160 Rio San Buenaventura (mythical), 89, 93, 97n23 riparian vegetation: and birds, 177; and exotics, 219; insects, 194; and water, 153, 154, 156, 157, 164n13, 221 Rivera, Juan Maria Antonio, 186n30 Rivera, Pedro de, 181 roads: and biological soil crusts, 228; and boundaries, 80; and energy extraction work, 162, 305, 341, 363; and environmental change, 129; Ferris-Bairoil Road, 30; gas field access roads, 50; and geology, 99n68; and Great Divide Basin, 115; hazards of, 78, 307, 308; and hydrology, 117; Interstate 80, 79, 80, 161, 163, 217, 262, 265n48, 305, 308; Lincoln Highway, 38, 260 –261, 262, 265nn37,44, 46; of Little Snake River valley, 313–314, 314–315n10, 315n11; network of, 77, 78 –79, 305; and plant diversity, 218, 222; Rawlins-Bairoil Road, ii; and sagebrush, 212, 213; and wildlife, 161–162 Robbins, Frank, 330, 333, 337nn25,27 Robinson, Lewis, 273 rock art, 79, 233, 245–250, 305, 359 394

Page 394

Rockefeller, John D., 364 Rock Springs, 85, 117, 128 –129 Rock Springs Grazing Association, 322, 333, 335, 337n25 Rock Springs Uplift, 94, 110, 112 –113, 138, 140 rock wren (Salpinctes obsoletus), 171, 173, 178 Rocky Mountain Fur Company, 355 Rocky Mountain glasswort (Salicornia rubra), 221 Rocky Mountain juniper ( Juniperus scopulorum), 221 Rocky Mountains, 153, 248, 249, 253, 256, 267 Rocky Mountain spotted tick, 263n13 rodeos, 333, 361n15 Rodríguez, Agustín, 185n3 Roosevelt, Franklin D., 350, 364– 365 Roosevelt, Theodore, 333 Rose Springs/East Gate points, 272 rough-legged hawk (Buteo lagopus), 173, 179 Routt National Forest, 115 rubber rabbitbrush (Ericameria nauseosa), 205, 206, 207, 210, 211 Rusling, James F., 258 –259, 264nn27,30 Russell, A. J., 129 Russell, Carl P., 124–125 Russell, Joellen, 96n8 Russell, Osborne, 127–128 Russian olive (Elaeagnus angustifolia), 218 Russian thistle (Salsola tragus), 157, 219, 222 Russin, Robert, 261

rusty lupine (Lupinus pusillus), 220 Ryan, B. T., 291n26

S Sackett, Delos B., 264n27 Sacramento River, 93, 99n84 sagebrush: adaptations of, 219; and birds, 153, 171, 176 –177, 179; characteristics of, 201– 205; chemistry of, 201–202; definition of, 201; and environmental change, 121, 124– 125, 128; and fires, 205, 206 – 207, 210, 211, 212, 213, 215n39; flowers and seeds of, 204– 205, 207, 208, 210; future of, 212 –213; and insects, 190, 192, 193, 194–195, 196, 197, 202, 204; leaves and water loss, 203–204, 219; and overgrazing, 212, 215n38; and Paleoindian period, 231; roots of, 204, 219; sagebrush steppe, 205–208, 305, 306; seedlings of, 205; Robert Louis Stevenson on, 260; and wildlife, 153, 154, 202, 208 sagebrush grasshopper (Melanoplus bowditchi), 194 sagebrush islands, 207 sagebrush lizard (Sceloporus graciosus), 147, 148, 149, 152, 153 sagebrush vole (Lemmiscus curtatus), 147, 152, 153 Sage Creek, 254 Sage Creek Basin, 14 Sage Creek Gap, 13 Sage Creek Ranch, 319 sage sparrow (Amphispiza belli), 147, 149, 153, 161, 171, 173, 177

Index

sage thrasher (Oreoscoptes montanus), 147, 153, 171, 173, 177 Saglauska, Doyle, 331 Saglauska, Jim, 331 Salicornia rubra (Rocky Mountain glasswort), 221 saline mineral reserves, 111 Salisbury, George, 324 Salish, 249 Salpinctes obsoletus (rock wren), 171, 173, 178 Salsola tragus (Russian thistle), 157, 219, 222 Salt Wells Creek, 117, 247 Salvelinus namaycush (lake trout), 146, 159 Sandberg bluegrass (Poa secunda), 206 Sand Creek Massacre, 293 Sand Dunes, 368 Sander vitreus (walleye), 147, 159 Sand Hills, 211 Sands, 366, 367, 370 sandstone: eolian sandstone, 92, 98n66; and Gosiute, 107; and Green River Formation, 100n107, 109; prevalence of, 88; and sagebrush, 208, 209; and terrain of Red Desert, 217; vermilion sandstone, 91 Sandy River, 125, 126, 127, 128, 134n66 Santa Fe Trail, 239, 255, 342 Sarcobatus vermiculatus (black greasewood), 206, 207, 210, 220, 224n18 Saskatoon, 89 savannah sparrow (Passerculus sandwichensis), 173, 177 Savery Creek Trail, 313, 315n11 Say’s phoebe (Sayornis saya), 171, 173, 178

Proulx_09_pp.377-400:Layout 1

8/4/08

scabies, 157, 161 scabland sagebrush (Artemisia rigida), 202 Sceloporus graciosus (sagebrush lizard), 147, 148, 149, 152, 153 Schell, Scott, 191 Schoenoplectus pungens (threesquare bulrush), 221 Schrum, Howard, 333 Scots immigrants, 307 Scott, Linda, 273 Scotty Lake, 179 scurvy, 240 seafloor spreading, 87 sedges, 153, 192, 196, 221 sediments: and environmental change, 121, 124; and geology, 88, 91, 92, 93–94, 95; and hydrology, 107–108, 109, 113, 114, 116, 117 seed-grinding evidence, 232, 233 Seedskadee rock art site, 246, 247 seep oil, 344 Selasphorus platycercus (broadtailed hummingbird), 172, 173, 174 selenium, 156 –157 self-defense, 355–356, 361n5 Seminoe Cutoff, 138 Seminoe Reservoir, 159 semipalmated sandpiper (Calidris pusilla), 173, 179 Separation Creek, 117 Sepsidae, 196 serviceberry (Amelanchier alnifolia), 121, 212, 240 Setophaga ruticilla (American redstart), 172, 173 71 ranch, 318, 319 shadscale saltbush (Atriplex confertifolia), 206, 220, 326 – 327n26

3:51 PM

Page 395

shale, 90, 107, 109 shamanism, 245, 248 Sharp ranch, 319 sharp-tailed grouse (Tympanuchus phasianellus), 147, 156, 222 Shaw, Al, 312 Shea, Pat, 335 sheep. See bighorn sheep; domestic sheep Sheep Creek, 137 Shepherd, William, 308 Sherman, William Tecumseh, 264n27, 286 Shimkin, D. B., 237n31 Shipley, F. W., 283 Shively, J. M., 240 short-eared owl (Asio flammeus), 173, 175 short-faced bear (Arctodus simus), 147, 155 shortite, 91, 98n58 Shoshonis: alliance with military, 275–276; and emigrants, 239 – 242, 242 –243n4, 243n16; and environmental change, 125, 127; and Fort Bridger, 235, 240 –241, 271, 272, 273, 274; and horses, 181, 182, 183, 184, 185n5, 186n26, 235, 237n33, 240, 241, 243n16; hunting and gathering of, 234; and Late Prehistoric period, 234; and Protohistoric period, 233, 234–235; and rock art, 246, 248, 249, 250; and South Pass, 275; and trade systems, 182, 183, 184, 234, 239 –242, 248 –249, 272, 273; Wilson’s report on, 237n33; and Wind River Reservation, 274 shrinking earth theory, 96n12 Shutterly, Lewis, 241

Sialia currucoides (mountain bluebird), 173, 178 Siberian craton, 91 Sierra Club, Wyoming chapter, 369 Sierra Madre, 212 Sierra Madre Forest Reserve, 318 Sierra Madre–Rawlins Uplift, 91 Sierra Nevada, 89 siltstone, 92, 107, 109 silver-haired bat (Lasionycteris noctivagans), 147, 152 silver sagebrush (Artemisia cana), 203, 204, 211, 219 Simmons, Marc, 186n15 Simon, Ben, 258 Simpson, Edward Hugh, 193 Simpson’s D, 193 Sioux: and Fort Halleck, 283; and Indian Road, 306; and Protohistoric period, 234; and Shoshonis, 275, 276; and Union Pacific Railroad, 293 Sioux War, 276, 286 site clusters, and rock art, 246 – 247 Sitta canadensis (red-breasted nuthatch), 173, 177 Sitta carolinensis (white-breasted nuthatch), 172, 174, 177 Skinner, Hugh, 240 Skull Point Gray Ware, 233 Skull Rim, 8, 10, 91, 305 Slater, Bill, 311 Slave craton, 89 small-footed myotis (Myotis ciliolabrum), 147, 152 Smith, Duane, 349 Smith, Gerald, 93, 99n80 Smith, Jedediah S., 89, 239, 242n3, 339, 340, 342 Smith, Joseph, 274

Index

Smith, Sidney, 127 smooth broome (Bromus inermis), 218 Snake River, 87, 93, 94, 99n84 Snake River Stockgrowers Association, 318 Snake-Sacramento river system, 93 snow deposition, 207, 211 snow fences, 153 snowmelt, 116 –117, 174, 220, 221 snowshoe hare (Lepus americanus), 147, 153 snowy owl (Bubo scandiacus), 173, 175 Socrates, 189 soda ash, 91, 98n59, 111, 114, 349 Soda Lake, 23 Soda Lake Flat, 43 sodic evaporites, 88, 97n20 sodium bicarbonate, 111 soil: biological soil crusts, 227– 228; and burrows, 207; and environmental change, 121, 123, 124, 126, 129 –130; and plant diversity, 219, 221; and soil water reservoir, 204, 205, 206, 207, 209, 210, 211 Soldierwolf, Mark, 371 Solvay Minerals Inc., 349 Sorbus (mountain ash), 273 Sorex monticolus (dusky shrew), 144, 145 South Pass: conservation of, 364, 369; and emigrants, 239; and environmental change, 125, 127, 128, 134n66; as feature, 305; and Fremont, 263n14; geology of, 97n20; and Shoshonis, 275; and trails, 343 South Pass City, 275 South Pass Formation, 98n65 395

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

South Pass Historic Landscape, 366, 368, 370, 372 South Pinnacles, 368 South Table Mountain, 94 Spanish-American War, 333 Spanish explorers/settlers, 155, 181–183, 184, 185nn3,4, 185–186n15, 329, 330 Spanish missions, 317, 329 Spanish Mustang Registry, 335 Spaulding, Payson W., 260 –261 Spea intermontana (Great Basin spadefoot), 144, 146, 151, 152 Spence, Mary Lee, 264n14 Spermophilus elegans (Wyoming ground squirrel), 147, 148, 151, 151, 207 Spermophilus lateralis (goldenmantled ground squirrel), 146 Spermophilus lateralis (gopher snake), 145, 146 Spermophilus lateralis ssp. wortmani (golden-mantled ground squirrel), 148 Spermophilus tridecemlineatus (thirteen-lined ground squirrel), 147 Spermophilus tridecemlineatus ssp. parvus (thirteen-lined ground squirrel), 148 spinies, 246, 249 spiny hopsage (Grayia spinosa), 205–206, 207, 220 Spizella breweri (Brewer’s sparrow), 145, 161, 171, 173, 177 Split Rock Formation, 98n65 Split Rock Gorge, 92 Sporobolus airoides (alkali sacaton), 221 spotted bat (Euderma maculatum), 147, 152 396

Page 396

spotted sandpiper (Actitis macularius), 173, 174, 176, 179 springs, 221, 318, 370 Stambaugh, Charles, 275 Stanford, Leland, 296 Stansbury, Howard, 79, 129, 134n66, 254–256, 262 –263n5, 344 States, Jack, 80 Steamboat Gap, geology of, 94, 95 Steamboat–Honeycomb Buttes, 91 Steamboat Mountain: and aspen, 222; and birds, 172, 174, 175, 176, 179; conservation of, 366, 367, 368, 370; and elk, 159; and environmental change, 126; geology of, 85, 94; and sagebrush, 209 Steamboat Rim, 209 Steele, Frederick, 291n17 Steen, Charlie, 350 Steidtmann, Jim, 91 Steller’s jay (Cyanocitta stelleri), 173, 180 Stevenson, Fanny Osbourne, 265n35 Stevenson, Robert Louis, 259 –260, 265n35 Steward, Julian H., 237n31 Stewart Creek horses, 330 stiletto flies, 195 Stillson, Jerome B., 289 stratified lake model, 108 Stratton, Tom, 325 streams: ephemeral, 116 –117, 118n32, 210; and insects, 194; meandering, 210 –211; perennial, 116; and plant diversity, 221; and ranching, 318; and sagebrush, 204, 210, 211; and wildlife, 152 –153, 156

striped skunk (Mephitis mephitis), 147, 160 stromatolites, 91, 98n54 Stuart, Robert, 125, 127 Stupich, Martin, 77 Sturges, David, 204, 207 Sturnella neglecta (western meadowlark), 172, 174 Sturnus vulgaris (European starling), 145, 160 St. Vrain’s Fort, 255 subalpine big sagebrush (Artemisia tridentata ssp. spiciformis), 202 subduction, 87– 89, 90, 91, 96n11, 97n35 Sublet (ghost town), 78 Sublette, Milton, 339 Sublette, William, 339, 342 Sublette Cutoff, 247 Sugarloaf rock art site, 247 Sunada, Edith, 303 Sun Dance, 250 Sundance Kid. See Longabaugh, Harry Sundin-Johnson Sheep Company, 325 Superior, Lake, 108 Superior craton, 89 Sutter, John A., 317 Swan, Alexander Hamilton, 319, 326nn14,15 Swan, Henry, 319 Swan, Tom, 319 Swan, William F., 319 Swan Land and Cattle Company, 319, 361n15 sweat bees, 195 sweep nets, 191, 192, 193, 196, 197 Sweetwater Canyon, 368 Sweetwater River, 95, 125, 210 –211, 221, 243n16

Index

Sweetwater River Preserve, 225n29 Sweetwater Rocks, 367 Swift, Jonathan, 195 swift fox (Vulpes velox), 144, 144, 145, 147 sylvatic plague, 149, 160 Symphoricarpos oreophilus ssp. utahensis (Utah snowberry), 211, 212

T Table Mountain, 209 Table Rock, 175, 178, 179 Tabuchi, Yoshio, 302, 304n35 Tachycineta bicolor (tree swallow), 174, 177 Tachycineta thalassina (violetgreen swallow), 174, 177 taconites, 88, 97n20 Tambora, 133n34 Tamiasciurus hudsonicus (red squirrel), 147, 153 Tanner, Russel L., 78, 249, 269n1, 340 –341 Taos (tribe), 182 Tate Geological Museum, 140 Taxidea taxus (American badger), 145, 148, 149, 152, 207 Taylor, William E., 241 Taylor Grazing Act of 1934, 160, 222, 364 telegraph, 156, 283–284, 285, 306, 356 temperatures: and Archaic period, 232; and plants, 221; and wildlife, 144, 145 terraces, geology of, 93 Tertiary period, 138 Tethyan orogen, 88, 96n8 Tethyan Sea, 88

Proulx_09_pp.377-400:Layout 1

8/4/08

Tetradymia canescens (gray horsebrush), 205 Texas, 267, 341 Texas longhorns, 253 Thaeler, C. S., Jr., 150 Thamnophis elegans (western terrestrial garter snake), 147, 148 Therevidae, 195 thickspike wheatgrass (Elymus lanceolatus ssp. lanceolatus), 206 thirteen-lined ground squirrel (Spermophilus tridecemlineatus), 147 thirteen-lined ground squirrel (Spermophilus tridecemlineatus ssp. parvus), 148 Thomas, Alfred, 185–186n15 Thomas, Craig, 368, 371 Thomas, David Hurst, 237n31 Thomomys clusius (Wyoming pocket gopher), 147, 148, 150, 150 Thomomys talpoides (northern pocket gopher), 146 Thomomys talpoides ssp. ocius (northern pocket gopher), 146 Thompson, Craig, 77, 367 Thornburgh, Thomas T., 286, 288 –289, 289, 291n17, 324 thoroughbraces, 259, 264–265n32 Three Mile Island accident, 351 three-square bulrush (Schoenoplectus pungens), 221 three-tip sagebrush (Artemisia tripartita ssp. tripartita), 202, 203 thrust loading, 94, 99n96 Thysanoptera (thrips), 193 Tibetan Plateau, 97n14 tick fever, 157 Tierney family, 324

3:51 PM

Page 397

tiger salamander (Ambystoma tigrinum), 147, 148 Timberlake, W. E., 320 timberline sagebrush (Artemisia rothrockii), 202 Tingey, W. M., 190, 193 titanotheres, 138, 139 Tokoi, Oskari, 346 –347 Tolar rock art site, 246, 248, 250 Tonaka, Yoshiye, 300 –301 topography: and maps, 254, 262 –263n5; and moisture, 128; and plant diversity, 218; and vegetation patterns, 207 Torbit, Steve, 367 totem figures, and rock art, 246 tourism, 217, 293, 339, 356, 361n8 Townsend, John Kirk, 127 Townsend Bill, 261 Townsend’s big-eared bat (Corynorhinus townsendii), 147, 152 trade systems: and Fort Bridger, 271, 272, 273; Prehistoric period, 187n40; and Shoshonis, 182, 183, 184, 234, 239 –242, 248 –249, 272, 273; and trappers, 311, 342 trails: conservation of, 364, 369, 370; and emigrants, 254–258, 259, 295, 306, 343; and freight wagon trains, 253; in Little Snake River valley, 313; and sheep, 317, 325n2; and U.S. Army, 283; and wildlife, 79, 156, 343. See also specific trails transcurrent ranges, 88 trappers: and Ashley, 339 –343, 352n1; of beavers, 124, 125– 128, 129, 134n71, 156, 305– 306, 340, 352n6; of bighorn

sheep, 160 –161; depictions of, 342, 352n11; emigrants as, 240, 241; and environmental change, 124, 125, 126, 127– 128, 134n71; and Fort Bridger, 271–272; in Little Snake River valley, 311, 314n1; Native Americans as, 125, 340, 342 –343, 344, 352n14, 355; and Shoshonis, 184, 239 Treaty of Fort Bridger (1868), 274 tree ring data, 121, 123, 124 tree swallow (Tachycineta bicolor), 174, 177 Trenholm, Virginia Cole, 234 Triad Land Company, 308 Trichoptera (caddis flies), 193 Trifolium pratense (red clover), 218 Trifolium repens (white clover), 218 Tringa flavipes (lesser yellowlegs), 173, 179 Tri-Territory Marker, 370 Troglodytes aedon (house wren), 172, 173, 177 trona: black trona water, 113; and geology, 91; and hydrology, 108, 110 –113, 112, 115; mining for, 349 –350, 363, 367 Tronquet, Gustave “Frenchy,” 357 trout, 113, 117. See also specific trout Trout Creek ranch, 318 Trout Unlimited, 113 tuberculosis, 157 Turdus migratorius (American robin), 172, 173, 177 Turkey Track ranch, 319 Twain, Mark, 127, 128, 134n66 Two Bar ranch, 318, 320 –321, 326n24, 357, 360, 361n15 Tympanuchus phasianellus (sharptailed grouse), 147, 156, 222

Index

U Udall, Stewart, 366 Uintah Utes, 234 Uinta Mountains: and environmental change, 121, 132n28; geology of, 90, 91, 97nn21,38; and horses, 182 Uinta Phase, 233 Uinta Phase points, 272 Uinta Range, 108, 112 uintatheres, 138 Uintatherium, 138 underfit valleys, 93, 99n87 ungulates, 153, 158, 161, 164n17. See also elk; mule deer; pronghorn Union Colony, 287 Union Pacific Coal Company, 275, 276 –279, 280n35, 345 Union Pacific Railroad: and arroyos, 129; Beadle’s coverage of, 294–296; and bison herds, 79; and Chinese workers, 297–299, 307; and coal mines, 343, 344, 345; effects on Wyoming of, 286, 293; and Fort Bridger, 275; and Fort Fred Steele, 286, 294; and grading, 294–295; and Greek workers, 300, 303; and Hispanic workers, 303; and Irish workers, 293, 297, 298, 303; and Japanese workers, 300 – 302, 307; labor force of, 293, 297–303, 307; and Morrow, 357; and Overland Trail, 309n6; and paleontology, 138; and plant diversity, 218; and wildlife, 156 Unruh, John, 240, 243n16 Upper Fremont Glacier, 123, 124, 133n34 397

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

uranium mining, 52, 62, 117, 348 –349, 350 uranium resources, 97–98n40, 343, 350 –351 urbanization, and plants, 222 Urosaurus ornatus (ornate tree lizard), 146 Ursus arctos (grizzly bear), 146, 157, 161, 163, 222 U.S. Army: and Bridger Pass, 255; and Chinese coal miners, 276, 278, 279, 345; and emigrants, 306, 309n8; and Fort Bridger, 272, 273, 274–275; and Fort Halleck, 284–286; and horses, 333–334; and Mexican-American War, 341. See also forts U.S. Department of Agriculture, 190 U.S. Department of Interior, 160, 368 –369 U.S. Forest Service, 318, 321, 322, 325nn5,6,7, 326n24, 334, 356 U.S. Geological Survey, 87, 262n2 U.S. Public Health Service, 117 U.S. Supreme Court, 355– 356 U.S. Topographical Corps, 256 Utah juniper (Juniperus osteosperma), 221 Utah snowberry (Symphoricarpos oreophilus ssp. utahensis), 211, 212 Ute Jack (Nicaagat), war chief, 287–288, 291nn19,25 Utes: and Fort Bridger, 271; and

398

Page 398

Fort Halleck, 284; and horses, 181–183, 185n2, 186nn19,30, 287; and Late Prehistoric period, 234; Charles S. Marsh on, 291n21; and Meeker Massacre, 286, 288 –290, 289, 291n20, 292nn29,31; and Mormons, 272; and Protohistoric period, 233, 234; relations with white settlers, 312; and rock art, 246, 248, 249; and White River Reservation, 286, 287–290. See also White River Utes

V Valles caldera, 94 Valverde, Antonio, 182 –183 vanadium, 350 Vanderburgh, W. H., 355, 361n1 Van Dyke, John C., 269n3 VanDyke, Kirk, 192 Vann, James, 256 –257 Vargas, Diego de, 182 varves, 109 Vasquez, Louis, 272, 273 vegetation, flora distinguished from, 214n1 vermilion sandstone, 91 Vermillion Creek, 109, 116 Vermivora celata (orange-crowned warbler), 146, 153, 173, 177 vesper sparrow (Pooecetes gramineus), 171, 174, 177 Vial, Pedro, 342 vibroseis, 340 –341 vigilante movements, 356 Villagrá, Gaspar Perez de, 185nn3– 4 Villasur, Pedro de, 183

violet-green swallow (Tachycineta thalassina), 174, 177 Vireo gilvus (warbling vireo), 172, 174, 177 Vireo vicinior (gray vireo), 144, 146 Virga, 5 Vivion, Charles, 360 Vlcek, David T., 249 volcanic arcs, 88, 89, 90 volcanic ash, 90, 91, 92, 98n40, 99n76, 111, 114, 139 volcanic eruptions, 133n34 volcaniclastic sediment, 91 volcanic sands, 109 Vulpes macrotis (kit fox), 144–145, 146 Vulpes velox (swift fox), 144, 144, 145, 147 Vulpes vulpes (red fox), 147, 160

W Wagon Bed Conglomerate, 98n64 Walker, Joseph Rutherford, 89, 239, 342 Wallace, Ernest, 182 walleye (Sander vitreus), 147, 159 Wamsutter: train derailment in, 42; skyline of, 32 Wamsutter Rim canneloid coal bed, 110 warbling vireo (Vireo gilvus), 172, 174, 177 Wardell, Thomas, 345 Wardell site, 233–234 Warner, Matt, 357–358, 359, 360, 361nn17,19, 362nn22,23 Warren, F. E., 278 Wasatch Formation, 9, 91, 140

Index

Washakie (Shoshone chief ), 249 Washakie Basin: Cherokee Rim, 40; geology of, 85, 91, 92; and Gosiute, 108; mud pot in, 70; and Paleoindian period, 231; paleontology of, 138, 139 Washington, Henry S., 100n97 wasps, 195 water: and arroyos, 124; and biological soil crusts, 227, 228; and birds, 152 –153, 171, 172, 174, 177, 179, 318; and Bitter Creek, 77; and coal-bed methane, 162; and emigrants’ trails, 256, 258, 259, 306; and hiking, 79; and ice age, 121; and livestock, 153, 160; and ranching, 318; and riparian vegetation, 153, 154, 156, 157, 164n13, 221; and sagebrush, 203, 204; and wildlife, 152 –153, 318. See also hydrology Waters, Frank, 352n7 watersheds, 116 waters of crystallization, 111 weathering: chemical weathering, 107, 108; physics of, 107 Weber, David J., 185n3 Wells Fargo, 309n6 Wentworth, Edward Norris, 317, 325n2 West, Neil E., 212 western chorus frog (Pseudacris triseriata), 147, 148 Western Heritage Alternative, 371 western hognose snake (Heterodon nasicus), 144, 147 western meadowlark (sturnella neglecta), 172, 174

Proulx_09_pp.377-400:Layout 1

8/4/08

Western Nuclear Corporation, 350 –351 western tanager (Piranga ludoviciana), 172, 174, 177 western terrestrial garter snake (Thamnophis elegans), 147, 148 Western Trails National Park, 364 western wheatgrass (Elymus smithii), 219, 220 western wood-peewee (Contopus sordidulus), 172, 174, 177 West Side Mining Company, 313 Westvaco Chlorine Products Company, 349 wetlands, 152 –153, 192, 196, 221 Wetzel, Curly, 333 Wheat, Carl, 254 White, Jack, 312 White, Richard, 356 white-breasted nuthatch (Sitta carolinensis), 172, 174, 177 white clover (Trifolium repens), 218 white-faced ibis (Plegadis chihi), 174, 179 Whitehorse Creek, 368 White Mountain: conservation of, 370; as feature, 305; geology of, 91; rock art of, 246, 248, 250 White River Agency, 288 White River Formation, 98n65 White River Reservation, 286, 287–290, 312 White River Utes, 234, 287– 288, 291n19, 292n31, 312, 314n5

3:51 PM

Page 399

white sucker (Catostomus commersonii), 147, 159 white-tailed jackrabbit (Lepus townsendii), 147, 148, 153 white-tailed prairie dog (Cynomys leucurus), 147, 148, 149, 149, 151, 160 Whitley, David, 245, 248 Wild Bunch, 358, 359, 361n19, 362n26 Wilderness Society Wyoming, 367, 369 –370 Wild Free-Roaming Horses and Burros Act of 1971, 160, 334, 336 Wild Horse Annie Act of 1959, 334 wild iris, 175 wildlife: adaptations of, 150 –152; and aspen, 222; challenges facing, 159 –163; and conservation, 157–159, 160, 161, 163, 166n65, 218, 223, 363–364, 366 –367, 369, 370, 371; and diseases, 160 –161; endemic vertebrates, 148, 149 –150, 151, 152; and energy extraction work, 163, 343, 363; exotic wildlife, 159 –160; influences of adjacent regions, 144–145, 144, 149; and invertebrates, 154–155; peripheral vertebrates, 144, 144, 149; physical setting and conditions, 143– 144; prehistory of, 155; recovery of, 157–159; and roads, 161–162; and sagebrush, 153, 154, 202, 208; and water, 152 – 153, 318; widespread vertebrates, 145, 148, 149. See also specific wildlife

Wilkins, James, 241 Wilkins Peak Member, 91 willet (Catoptrophorus semipalmata), 174, 174 Williams, Joseph, 239 Williams, Walt, 333 Willow Creek Ranch, 318 Willow Hill, 4 willows: and birds, 171, 177, 178; and environmental change, 125, 126, 128; and overgrazing, 156; range of, 221; and water, 153, 154 Wilson, Gilbert, 331 Wilson, John, 237n33 Wilson, W. W. “Wiff,” 320, 321, 331, 347–349, 353n21, 359 – 360 Wilsonia pusilla (Wilson’s warbler), 174, 178 Wilson’s phalarope (Phalaropus tricolor), 174, 174 Wilson’s warbler (Wilsonia pusilla), 174, 178 Wind River, 88, 123, 128 Wind River Basin, 109, 212 Wind River fault, 90 Wind River–Granite Mountains foreland uplift, 90, 91 Wind River Mountains: and environmental change, 121, 123, 127; geology of, 89, 95, 97n21; plants of, 217; Preuss on, 265n35 Wind River Range, 94, 108, 113, 124, 221 Wind River Reservation, 274, 370 Wind River thrust, 94 winds: and aridity, 144, 210; and biological soil crusts, 227, 228; and plants, 220; and road haz-

Index

ards, 308; and snow distribution, 207, 211 winterfat (Krascheninnikovia lanata), 220, 223, 224n17 Wise, Tom, 313 Wister, Owen, 361n9 wolverines, 222 women: employment during World War II, 302; population of, 308; and trappers, 342 – 343, 352n14 Wood, Raymond, 184 Woodruff site, 234 Worcester, Don, 329 World War I, 333 World War II, 301–302, 334 Wyomide, 91 Wyoming Basin: and beaver, 125; and environmental change, 121, 124, 126, 128; and horses, 181–184 Wyoming big sagebrush (Artemisia tridentata ssp. wyomingensis): and basin big sagebrush, 209, 210; and Jack Morrow Hills, 212; leaves of, 203, 204, 211; prevalence of, 202, 219; and sagebrush steppe, 205– 208 Wyoming craton, 89 –91 Wyoming Game and Fish Department, 159, 162, 194, 209, 366 –367 Wyoming Grasshopper Information System, 190 Wyoming ground squirrel (Spermophilus elegans), 147, 148, 151, 151, 207 wyomingites, 88, 97n20 Wyoming Outdoor Council, 79, 363, 366, 369, 370

399

Proulx_09_pp.377-400:Layout 1

8/4/08

3:51 PM

Wyoming pocket gopher (Thomomys clusius), 147, 148, 150, 150 Wyoming Range, 108 Wyoming salient, 98n43 Wyoming State Historic Preservation Office, 254 Wyoming State Penitentiary, 359, 360 Wyoming Stock Growers Association, 319 Wyoming Territorial Penitentiary, 262n4 Wyoming three-tip sagebrush

400

Page 400

(Artemisia tripartita ssp. rupicola), 202, 208 Wyoming Trona Company, 113 Wyoming Wilderness Association, 368 Wyoming Wildlife Federation, 367 Wyoming Wool Growers Association, 322

X Xyrauchen texanus (razorback sucker), 147, 159

Y Yamparika Comanches, 182, 186n17, 234 Yampa River, 7, 93, 116 Yancey ranch, 318 yellow-rumped warbler (Dendroica coronata), 172, 174, 177 yellow sticky traps, 191, 192, 197 Yellowstone: and environmental change, 123; geology of, 92, 94; hydrology of, 109; plants of, 217

Index

Yellowstone caldera, 94 Yellowstone National Park, 157, 158, 159, 201, 218 Yellowstone Sheep, 322 yellow warbler (Dendroica petechia), 174, 177 Yosemite Valley, 89 Young, Brigham, 272, 274 Young, James A., 212 Young Warriors Society, 370

Z Zhu, Li Ping, 298