Sie sind auf Seite 1von 417

Mountains

and Plains

Dennis H. Knight
George P. Jones
William A. Reiners
William H. Romme

Mountains
and
Plains

The Ecology
of Wyoming
Landscapes
S e c o n d Ed i t i o n

New Haven and London

University of Wyoming
Laramie, Wyoming
www.uwyo.edu/biodiversity

Published with assistance from the University of Wyoming

Library of Congress Cataloging-in-Publication Data

Biodiversity Institute (wyomingbiodiversity.org).

Knight, Dennis H.

Published with assistance from the Mary Cady Tew Memorial


Fund.

Mountains and plains: the ecology of Wyoming landscapes/


Dennis H. Knight, George P. Jones, William A. Reiners, William H.
Romme.Second edition.

First edition 1994. Second edition 2014. Copyright 1994, 2014

pagescm

by Yale University. All rights reserved. This book may not be

Includes bibliographical references and index.

reproduced, in whole or in part, including illustrations, in any

ISBN 978-0-300-18592-8 (pbk. with flaps : alk. paper)

form (beyond that copying permitted by Sections 107 and 108

1.EcologyWyoming. 2.Landscape ecologyWyoming.

of the U.S. Copyright Law and except by reviewers for the public

I.Title.

press), without written permission from the publishers.

QH105.W8K582014
577.09787dc23 2014015382

Yale University Press books may be purchased in quantity for


educational, business, or promotional use. For information, please
e-mail sales.press@yale.edu (U.S. office) or sales@yaleup.co.uk
(U.K. office).
Designed by Nancy Ovedovitz. Set in Stone Serif Bold and
Beton Extra Bold display by Princeton Editorial Associates Inc.,
Scottsdale, Arizona.
Printed in China through World Print Ltd.

A catalogue record for this book is available from the British


Library.
This paper meets the requirements of ANSI/NISO Z39.48-1992
(Permanence of Paper).
10987654321
Frontispiece photo of Gannett Peak and Dinwoody Creek by Scott
Copeland (www.scottcopelandimages.com); see fig. 2.5.

To our students and colleagues,


from whom we have learned so much

Nature is an open book. . . . Each grass-covered hillside


is a page on which is written the history of the past,
conditions of the present, and predictions of the future.
. . . Let us look closely and understandingly, and act
wisely, and in time bring our methods of land use and
conservation activities into close harmony with the
dictates of nature.
John E. Weaver, 1954

Contents

Preface, ix

Part One. Wyoming and

Part Four. Foothills and

Acknowledgments, xi

the Rocky Mountain West

Mountains

1. Introduction, 3

10. Escarpments and the

2. Landscape History, 12
3. P resent-Day
Environments and
Climate Change, 27

Foothill Transition, 155


11. Mountain Forests, 173
12. The Forest Ecosystem,
200
13. M
 ountain Meadows
and Snowglades, 221

Part Two. Wetlands


4. R iparian Landscapes, 45

14. Upper Treeline and


Alpine Tundra, 230

5. Marshes, Playa
Wetlands, Wet
Meadows, and Fens, 66

Part Five. Landscapes of


Special Interest
15. The Greater

Part Three. Plains and


Intermountain Basins
6. Grasslands, 83
7. Sagebrush, 109
8. Desert Shrublands and
Playas, 131

Yellowstone
Ecosystem, 245
16. The Black Hills, Bear
Lodge Mountains, and
Devils Tower, 266
17. The Laramie Basin, 282

9. Sand Dunes, Badlands,


Mud Springs, and
Mima Mounds, 142

Part Six. Sustainable Land


Management
18. Using Western
Landscapes, 305

Epilogue, 319

Appendix A: Conversion
Table, 321
Appendix B:
Characteristic Soil
Types, 322
Notes, 325
References, 349
Index, 387

This page intentionally left blank

Preface

Many changes have occurred in the Rocky Mountain

lands, forests, subalpine meadows, and alpine tundra.

region since 1994, when the first edition of this book

The ecology and management of croplands are beyond

was published. Wildlife habitat has been fragmented

the scope of this book, but we discuss the ecological

at alarming rates, the once abundant sage-grouse has

effects of irrigation, invasive species, livestock, feral

been proposed for protection by the Endangered Species

horses, elk, prairie dogs, and wolves. The book ends

Act, invasive plants have become more common,

with a chapter on current issues pertaining to land

wolves have been reintroduced, climate change is now

management and conservation.

well documented, epidemics of forest insects are more

Plant ecology is emphasized, because vegetation pro-

widespread, forest fires are more frequent, and new ap-

vides habitat and gives character to the landscape, but

proaches for conservation have been adopted. Our goal

we also discuss plant-animal interactions, hydrology,

for the second edition, 20 years later, is to provide a new

nutrient cycling, and the effects of disturbances, such

synthesis of the ecological research that is pertinent to

as fire, insect epidemics, and timber harvesting. Most

natural resource management, with a focus on the ecol-

chapters consider the implications of climate change.

ogy of Wyoming and adjacent parts of Colorado, Utah,

The chapters vary in length and structure, based on the

Idaho, Montana, South Dakota, and Nebraska.

information available and our perception of what most

The book has six parts. The first is an introduction


to ecology, conservation biology, and the concept of

readers will find interesting. We provide guidance to


further information in the endnotes.

ecosystem services. It also includes a chapter on the

Recognizing the diverse backgrounds of people inter-

geological development of the diverse landscapes of the

ested in western landscapes, we have written this book

region and changes during the past several thousand

for anyone intrigued by the natural history and wildlife

years, plus a chapter on present-day environments and

of the region. The text is essentially free of technical

climate change during the past few decades. The parts

terms, but those few that remain are defined where they

that follow focus on riparian and nonriparian wetlands,

first appear. Definitions can be located using the index.

semi-arid basins and plains, foothills and mountains,

We have used common names for plants and animals

and three landscapes of special interestthe Greater

throughout the book, but Latin names for most species

Yellowstone Ecosystem; the Black Hills, Bear Lodge

are provided in the tables. English units of measure are

Mountains, and Devils Tower; and the Laramie Basin.

used almost exclusively, with a metric conversion table

All the major kinds of ecosystems are discussed: flood-

in appendix A.

plains, marshes, fens, mixed-grass prairie, sagebrush

Online resources are now of great value for ecologists

steppe, desert shrublands, sand dunes, badlands, wood-

and land managers, providing quick access to additional

ix

Preface

information, photos, and distribution maps for the

about. We also have been involved with the application

plants, animals, and virtually every topic mentioned

of our research and that of other scientists to perplex-

in the text. Some of the pertinent website addresses

ing challenges faced by society today. Discussions with

are in the endnotes. Our book has a website as well. If

agency and private land managers have broadened our

you learn something that we should know about, go

perspectives. Our goal in writing this book is to present

to mountainsandplains.net. There you can make com-

information that is pertinent to ongoing debates and

ments, ask questions, read our responses to questions

that illustrates the long history of a place that has been

already submitted, and see additional photos. Other

our home. Two important questions are: How can we

useful links are at http://www.uwyo.edu/biodiversity/,

use ecosystems in ways that enable future generations

the website for the Biodiversity Institute at the Univer-

to benefit from them as well? And how can we antic-

sity of Wyoming.

ipate and adapt to changes associated with a warmer

We have worked together for more than 30 years.

and drier climate while conserving biological diversity?

Our careers have blended education at various levels

Answers require policies informed by science but cogni-

with research on many of the ecosystems that we write

zant of diverse values and traditions.

Acknowledgments

We have benefited greatly from the encouragement,

John Hauer, and Hollis Marriott. Helpful suggestions for

knowledge, and talents of many colleagues during the

the Laramie Basin chapter were provided by Don Boyd,

development of this book. Ken Driese provided numer-

Tony Hoch, and Larry Munn; Martin Curry, Jerry Han-

ous photographs and some of our maps, and he, Dan

sen, Bern Hinckley, Mindy Meade, Ruben Vasquez, Kim

Binkley, and Phil White offered detailed, helpful sug-

Viner, and Lindsay Wheat answered questions about

gestions on all of the chapters. Phil was extraordinarily

the Basin. The chapter on land management issues was

diligent in assisting with library work and last-minute

improved with the suggestions of Indy Burke, Brian

details. Chris Nicholson contributed the climate maps

Kuehl, Jay Lillegraven, and Chris Madson.

in chapter 3 and Ramesh Sivanpillai created three of the

Others have helped as well. Especially important

maps in chapter 17. Their work, along with that of vari-

has been the research of our students over the years.

ous photographers, is identified in the figure captions.

They have greatly influenced how we think about ecol-

Many colleagues offered helpful suggestions on early

ogy and natural resource management. Numerous col-

drafts of various chapters. Don Boyd, Steve Buskirk,

leagues and agency personnel provided information,

Steve Jackson, Jay Lillegraven, Larry Munn, Bryan Shu-

including Ken Anderson, Bob Dorn, Bonnie Heidel, Jeff

man, Art Snoke, and Cathy Whitlock assisted with our

Lockwood, Jim Lovvorn, Chris Madson, Jill Morrison,

discussion of geologic history, modern-day environ-

Larry Munn, Ginger Paige, Scott Schell, Pete Stahl, Alan

ments, human history, and climate change. The for-

Ver Ploeg, Danny Walker, and Lee Whittlesey. Ron Hart-

est chapters were improved by the comments of Craig

man, Bonnie Heidel, and Ernie Nelson assisted with

Benkman, Peter Brown, Brent Ewers, Tim Fahey, Dan

plant identification, and Steve Buskirk answered ques-

Tinker, and Tom Veblen, and the grassland and shrub

tions about Wyomings fauna. Allory Deiss drew essen-

land chapters by the comments of Jeffrey Beck, Angela

tially all the line figures, Judy Knight contributed four

Hild, Alan Knapp, William Lauenroth, Brian Mealor,

illustrations, and Joy Handley assisted with the tables

Rachel Mealor, and Elise Pendall. The wetland chapters

and plant nomenclature. Jean Thomson Black at Yale

benefited from the comments of David Cooper, Laurie

University Press was very supportive, as she was with

Gilli
gan, Bonnie Heidel, and Joanna Lemly, and the

the first edition. Her associates Samantha Ostrowski

alpine meadow and tundra discussions from suggestions

and Mary Pasti, along with Peter Strupp and Cyd West-

by Tim Seastedt and William K. Smith. The chapter on

moreland of Princeton Editorial Associates, drew our at-

the Greater Yellowstone Ecosystem was improved with

tention to details and problems that we had overlooked.

the help of Andrew Hansen, Matt Kaufmann, and Paul

Nancy Ovedovitz designed the book; Christy Spielman

Schullery, and the Black Hills chapter by Beth Burkhart,

and Bryce Tugwell developed our website.

xi

xii

Acknowledgments

Over the years our research has been made possible

Wyoming Depart
ment of Agriculture, the Wyoming

by grants from the National Science Foundation, the

Game and Fish Department, and the U.S. Forest Service.

U.S. Forest Service, the U.S. Department of Energy, the

For the second edition, we received a grant from the

U.S. Department of Agriculture, the Bureau of Land

Biodiversity Instit ute at the University of Wyoming, di-

Manage
ment, the University of WyomingNational

rected by Carlos Martinez del Rio. Thanks to his encour-

Park Service Research Center, the Wyoming Water Re-

agement and support, along with that of Indy Burke,

sources Research Institute, The Nature Conservancy,

director of the universitys Haub School of Environ-

and the Mellon Foundation. Grants to facilitate the

ment and Natural Resources, this book is much better

publication of the first edition were provided by the

than it would have been otherwise.

Part One Wyoming and the

Rocky Mountain West

This page intentionally left blank

Chapter 1

Introduction

Straddling the Continental Divide and with an aver-

grasslandsthe Thunder Basin National Grassland in

age elevation of 6,700 feet above sea level, the land-

northeastern Wyoming and the Pawnee National Grass-

scapes of Wyoming are similar to those encountered

land in northern Colorado. Only about 4 percent of

by explorers in the early 1800s. There are now roads,

Wyoming is currently under cultivation, leaving large

ranches, farms, cities, and industrial developments,

tracts of land that have never been plowed and are still

but outdoor enthusiasts, naturalists, and scientists

dominated by native plants.1 Overall, the most signifi-

are attracted by the wildlife and natural ecosystems

cant landscape changes have occurred in three set-

that still exist over large areas. The persistence of such

tings: (1) along rivers at low elevations, where irrigation

amenities can be attributed to a cool climate, rugged

water is available and people enjoy having their homes;

terrain, and extraordinary features that are highly

(2)in the foothills, where many homes have been con-

valued. Yellowstone National Park was established in

structed and fires have been suppressed; and (3) where

1872the worlds first national park. Nineteen years

the extraction of energy resources is feasible.

later, in 1891, a forest reserve was established nearby.

During the past 20 years, Wyoming and adjacent

It became the Shoshone National Forest. In 1906 the

states have experienced surprisingly rapid changes.

nations first national monument was established at

Still, there are many opportunities for learning about

Devils Tower in northeastern Wyoming, and Grand

the natural history of the Rocky Mountains and western

Teton National Park was established in 1929. Since

Great Plains. Bison have been replaced in most areas by

then fifteen wildlands in Wyoming have been desig-

cattle, but pronghorn, elk, and deer are still common.

nated Wilderness by Congress. National parks, wilder

Livestock have grazed some rangelands heavily, but

ness, and wilderness study areas account for about

bison may have done the same. Grizzly bears, elk, and

9percent of the states land (figs. 1.11.5).

wolves no longer wander across the basins and plains

In addition, native prairie and shrublands persist in

to the extent they once did, but all are found in some

some areas, primarily because 37 percent of the state is

areas.2 Roads, timber harvesting, and industrial devel-

above 7,000 feet elevation, where growing seasons are

opments have fragmented mountain forests and basin

too short and cool for cropland agriculture. At lower

shrublands, but not everywhere. The question is: How

elevations water is more likely the limiting factor, with

can the land be managed so that highly valued natu-

most agriculture requiring irrigation. Farmers tried vari-

ral resources are used wisely during a time of alarming

ous crops in the late 1800s, but their fields commonly

rates of habitat degradation and climate change? The

failed. The federal government purchased abandoned

sciences of ecology and conservation biology are perti-

homesteads in the late 1930s, and some became national

nent to answering this question.

4 Wyoming and the Rocky Mountain West

Fig. 1.1.Elevation map of Wyoming and neighboring parts of


Montana, South Dakota, Nebraska, Colorado, Utah, and Idaho.
The lowest elevations in the state are in the northeast. The
names of physical and cultural features are shown in figs. 1.2
and 1.4. Land cover is illustrated in fig. 1.5, and major rivers

are identified in fig. 4.1. The total area of Wyoming, including that portion of Yellowstone National Park in the state, is
97,814 square miles. The state is bounded by longitudes 1043
and 1113 and latitudes 41 and 45. Cartography by Ken
Driese.

Ecology

ing the interactions among plants, animals, micro-

Most ecologists would argue that it is important to


protect air and water quality, conserve biological
diversity, and maintain soil productivity. Yet advocacy
for env ironmental protection is not ecology per se.
Rather, ecology is a science dedicated to understand

organisms, humans, and their environment. The


interactions are studied to be understood, not to be
judged as good or bad. Ecologists strive to be objective
through systematic measurements, experimentation,
and analysis. Many also work to ensure that decision-

Introduction

LL
PL OW
AT S
EA TO
U NE

YE

O N MTS.

KIE

MT
S.

OW

LC

SP

CH

AR

RI

ER

D
IN

GRA
N

RAWLINS
UPLIFT

SHIRLEY
BASIN

HANNA
BASIN

LARAMIE
BASIN

DENVERJULESBURG

ME

BASIN

NORTH
PARK

BO
NE .
CI
DI MTS

TS.

UINTA M

.
MTS

BASIN
ROCK
SPRINGS
UPLIFT

RA
ER RE
SI AD
M

SS

MT
S.

DIVIDE

WASHAKIE
BASIN

FO

ITE

E
MI
RA
LA

OVERTHRUST BE
LT

TS.

GREAT

RIVER

IL SY
NCLINE

KM

S.

GREEN

BASIN

REE

WIND
RIVER
BASIN

MT

RIVER
BASIN

CA

ER

POWDER

FT
LI
UP

TE
T

TS

S
LL
HI

BASIN

K
AC
BL

BIGHORN

KA

VE GR
O
M NTR S
TS E
.

TS.
RN M
HO
IG
B

O
ABSA R

SH
WA
N
KSO
JAC OLE
H

WASATCH MTS.

PRYOR
MTS.

BEARTOO
TH
MT
S.

HAR
T
UPL VILLE
IFT

MADISONGALLATIN
MTS.

100 km

100 mi

Fig. 1.2. Mountains and basins illustrated in the previous


figure. See Love and Christensen (1985) for a detailed geologic
map of Wyoming. Adapted from Blackstone (1988) and Mears
(1993).

makers are provided with the best ecological informa-

to understand how organisms are adapted to their en

tion available.

vironment, whereas population ecologists seek to under-

The goal of modern ecological research, as in other

stand why population sizes fluctuate through time and

sciences, is to understand patterns and processes in

vary from place to place. Community ecologists focus

nature well enough to make predictions. This includes

on the interactions among coexisting species in a spe-

anticipating the effects of environmental changes on

cific areathe community. Others study ecosystems,

the distribution and population sizes of all kinds of life,

defined as areas where plants, animals, and microbial

and on the movement of water, nutrients, and energy.

organisms interact with one another and their environ-

Theoretical ecologists search for generalizations that

ment. Ecosystem ecologists commonly measure flows

apply everywhere. Others are empirical; that is, they

of energy, water, and nutrients through the soil, atmo-

focus on specific areas and address interactions asso-

sphere, detritus, and living organisms (fig. 1.6). Often

ciated with, for example, specific tracts of forests and

the innumerable species are grouped into green plants

rangelands. Each approach is beneficial for the other.

(the producers) that convert the energy of sunlight into

Modern ecology can be divided into numerous sub-

new plant material, herbivores that eat plants, carnivores

disciplines. To illustrate, physiological ecologists work

that consume other animals, omnivores that depend on

6 Wyoming and the Rocky Mountain West

Fig. 1.3. Land jurisdiction in Wyoming and adjacent parts of


neighboring states. Federal and private lands make up 48 and
43 percent of Wyoming, respectively. Most federal land is administered by three agencies: the U.S. Forest Service (national
forests and national grasslands), the National Park Service
(parks, monuments, and recreation areas), and the Bureau of

Land Management. Approximately 5 percent of the state has


been designated by Congress as Wilderness, all on federal
land. State and tribal lands cover 6 and 3 percent of the state,
respectively. State-owned lands are widely distributed in
small tracts and are not separated from private land on this
map. County lines are shown. Cartography by Ken Driese.

both food sources, and detritivores and decomposers that

how the abundance of different species and their influ-

derive their energy and nutrients from dead plants and

ences are affected by varying environmental conditions,

animals. As the populations of a community or ecosys-

and how they change with time following disturbances.3

tem change, other attributes change as well (for example,

Some ecologists specialize in the study of plants, whereas

rates of streamflow, photosynthesis, herbivory, and the

others concentrate on bacteria, fungi, insects, birds,

cycling of nutrients). Of interest to many ecologists is

mammals, or other organisms.

Introduction

Fig. 1.4. Highways, cities, and larger towns in Wyoming. The


two largest cities are Cheyenne and Casper, with populations
in 2012 of 62,000 and 58,000, respectively. Cartography by
Ken Driese.

Another branch of ecology is landscape ecology. Ecol-

objectives of a study or management challenge. The area

ogists working at this scale commonly focus on large het-

included may be as small as a pond or as large as several

erogeneous areas, for example, square miles of grasslands

counties or states, or even continents.

and shrublands. They are curious about how the land-

Ecology is a diverse science that fosters an improved

scape has changed in recent decades or centuries, and

understanding of all the various forms of life in an area

how it might change in the future. Another goal is to

and the ecosystems they inhabit. To facilitate ecologists

determine the causes and effects of different landscape

work, diagrams are drawn to illustrate the plethora of

patterns. Aerial photographs, satellite images, and maps

possible interactions, such as those illustrated in fig.1.6.

are important tools for their work. The terms landscape

Some of the drawings are similar to those prepared by

and ecosystem are sometimes used interchangeably. To

engineers, but there is a significant differencenatural

illustrate, the Yellowstone landscape made famous by the

ecosystems were not designed and constructed by

artist Thomas Moran in the 1870s is a mosaic of commu-

humankind. Rather, they emerged after millions of

nities and is now often referred to as the Greater Yellow-

years of evolution and centuries of ecosystem develop-

stone Ecosystem (see chapter 15). The boundaries of an

ment. Much remains to be learned about the species

ecosystem or landscape are established according to the

and processes that enable their long-term persistence.

Introduction

SOLAR RADIATION

Transpiration
Atmosphere

Photosynthesis
Heat

Herbivores
Leaves

Evaporation
to atmosphere

Carnivores

Fruit
Seeds

Omnivores

Rain
Snow
Surface
runoff

Nitrogen
fixation

Stems
Detritus

Roots
Mycorrhizae

Herbivores

Detritivores

Soil surface

Carnivores
Soil organic
matter

Soil
solution

Omnivores

Decomposers
Subsurface
runoff

Mineral
soil

Weathering

Fig. 1.6. Major components (indicated by boxes) and interactions (arrows) of a terrestrial ecosystem. Arrow width indicates
the relative amount of energy or water moving along a pathway. Temperature, water and nutrient availability, and growing season length determine the rates of transfer between
components. The irregular shapes indicate sources of water
and nutrients. Complex food webs exist above and below the
soil surface, both of which are linked by the organic matter
on the soil, known as detritus, litter, or mulch. Such simple
diagrams do not convey the complexity caused by the diverse
group of organisms represented by each box. Changes in one
component or process cause changes in others.

Conservation Biology
As for ecologists, conservation biologists can take several approaches to their research. In general, they are
experts on rarity and what, if anything, can be done
about it. Some species are rare because their habitat has
been degraded by human activity; others are naturally
rare and are found only in one or a few small areas. Ultimately, conservation biologists work to facilitate sound
management programs that lead to the maintenance or
recovery of threatened species, thereby reducing the need
for the strict mandates of the Endangered Species Act. In

Fig. 1.5. (left) Major vegetation types plus cultivated


land, lakes, reservoirs, and urban and industrial developments.Sagebrush-dominated shrubland is most widespread
(33 percent of the land area), followed by mixed-grass prairie
(18 percent) and lodgepole pine forest (7 percent). Some
categories are not shown, because they occur in patches too
small for the scale of this map. These include most riparian
woodlands, shrublands, and meadows; most subalpine meadows; most aspen groves; small woodlands of ponderosa pine
and limber pine; woody draws; playas with greasewood; small
cultivated fields; white spruce groves in the Black Hills; and
foothill grasslands and shrublands.Subheadings indicate associated species in the foothills (fh), the Greater Yellowstone
Ecosystem (gye), the Black Hills (bh), and the Sierra Madre
(sm).Adapted from the national land cover map (2011) of the
U.S. Geological Survey Gap Analysis Program. Land cover
percentages are from Driese et al. (1997). Cartography by Ken
Driese.

Wyoming, seven plants, six mammals, six birds, six fish,


and one amphibian are currently protected by federal law
or are under consideration for protection (figs. 1.71.9).
Others are of special interest because they are endemics, that is, they are found nowhere else.5 Such species
along with those that are common and widespread, and
all the varieties in each speciescompose the biological
diversity of an area.
The conservation of rare and endangered species
requires information gained by scientists in ecology as
well as other disciplines. More specifically, biologists
with expertise in evolution and species identification are
needed to verify that so-called rare species are verifiable
species and are indeed rare. Some of them are difficult to

10

Wyoming and the Rocky Mountain West

distinguish from related species that are common, which


leads to controversy about whether the restrictions of the
Endangered Species Act are needed. Do the advocates of
such action really know? The expertise of systematists,
taxonomists, and geneticists is required. Its clear, however, that scientists cannot accomplish the goal of conservation by themselves. Stemming the tide of habitat
destruction and extinction requires joining forces with
people from all walks of lifelandowners; policymakers;
nongovernmental organizations; and local, state, and
federal agencies. Collaborating with so many parties is
complicated but necessary as increased value is placed on
rare species and other limited resources. Notably, some
resources, though not traditional commodities, have the
potential of providing ecosystem services that can be
bought and sold.

Ecosystem Services
Earths ecosystems provide many benefits in addition
to habitat for wildlife and rare species. These benefits,
often referred to as ecosystem services, include

Fig. 1.7.The black-footed ferret (Mustela nigripes), a member of


the weasel family, was believed to be extinct until a few dozen
individuals were found in 1981 on a ranch near Meteetsee,
Wyoming. With the collaboration of private landowners,
biologists, and state and federal agencies, there are now about
1,000 ferrets living in twenty large prairie dog colonies in
Arizona, Colorado, Kansas, Montana, New Mexico, South
Dakota, Utah, Wyoming, Canada, and Mexico. Adults are
1824 inches long. The species is protected by the Endangered
Species Act. Photographed in Arizona by George Andrejko,
Arizona Game and Fish Department.

Fig. 1.8.The worldwide distribution of desert yellowhead


(Yermo xanthocephalus) is confined to several hundred acres
of desert shrubland in central Wyoming. A member of the
sunflower family, this plant was first discovered and named
by Robert Dorn in 1990 and is protected by the Endangered
Species Act. It grows to a height of about 12 inches. Photo by
Bonnie Heidel.

erosion control,
maintenance of soil fertility,
inhibition of invasive plants,
provision of clean water,
air filtration,
mitigation of droughts and floods,
provision of habitat for insects that pollinate crops,
detoxification and decomposition of waste materials,
sequestration of carbon dioxide,
control of agricultural pests, and
provision of meat from areas that are too dry to cultivate using standard agricultural practices.
Such benefits are provided by the diversity of native
plants, animals, and microbes that thrive in the harsh
environments of the region.

Introduction

fossil fuels, the value of ecosystem services is estimated


at trillions of dollars per year.6 The Farm Bill of 2008,
passed by Congress, called for measuring the environmental benefits gained from managing farm and ranch
lands more carefully.
Considering their value, it should not be surprising
that some ecological services now can be bought and
sold, thereby creating opportunities for landowners,
businesses, consumers, and governments to become
more directly involved with supporting sustainable
resource management. An example is the buying and
selling of carbon credits as a means of reducing carbon dioxide emissions to the atmosphere, a major
Fig. 1.9.The Wyoming toad (Anaxyrus baxteri), less than 2.5
inches long, is found nowhere else except in an area of about
2 square miles in the Laramie Basin. It has been raised successfully in captivity, but there are currently only thirty to
a hundred individuals living in the toads native habitat. It
is protected by the Endangered Species Act. Photo by Sarah
Armstrong.

contributor to global climate change. Similarly, the


Conservation Reserve Program of the U.S. Department
of Agriculture provides monetary incentives for farmers to establish perennial plant cover on erodible land,
thereby enabling carbon sequestration as well as soil
stabilization and water purificationpowered by solar
energy rather than fossil fuels. Other examples are discussed in chapter 18.

Ecosystem services generally have been undervalued


or overlooked, leading to inevitable and undesirable

In sum, whereas ecologists study adaptations, popu-

consequences. Their value becomes clear when the costs

lations, communities, and ecosystems, conservation

are summed for the time, labor, and energy required for

biologists strive to develop strategies for conserving our

water purification, herbicides, insecticides, fertilizer,

evolutionary heritage. And now, a newly emerging mul-

and the control of soil erosion. When economists cal-

tidisciplinary group of professionals seeks to identify

culate all the costs of replacing such services with mod-

and profit from investments in maintaining ecosystem

ern technology, much of which requires the burning of

services.

11

Chapter 2

Landscape History

Millions of Years Ago

ERA

PERIOD

EPOCH
Holocene

Rocks exposed in canyons, escarpments, and moun-

Quaternary

tains provide the data geologists need to interpret the

2.6
Pliocene

climate was tropical. The water receded and advanced


many times, depositing sand in some places and more
finely textured silt and clay elsewhere. Sandstones, silt-

CENOZOIC
(Age of Mammals)

changes have occurred. About 350 million years ago,


become Wyoming was covered by seawater and the

stones, claystones, and other kinds of sedimentary rock

Miocene
23
Oligocene
34

Paleocene

limestones. An example is the Madison Limestone that


MESOZOIC
(Age of
Reptiles)

forms many escarpments in the region today.

dragonflies with 30-inch wingspans. Primitive vascular


plants were abundant, including tree-sized clubmosses,
ferns, and horsetails. Fossilization was less frequent on

Fig. 2.1. Geologic time chart showing the ages (millions of


years ago) of eras, periods, and epochs. The Paleogene and
Neogene periods of the Cenozoic were previously known as
the Tertiary. Adapted from the International Commission on
Stratigraphy (http://www.stratigraphy.org).

12

PALEOZOIC
(Age of Fishes)

ginkgo, pine, spruce, fir, and sequoia. Mountain ranges

Jurassic
Triassic

Pennsylvanian
Mississippian
Devonian
Silurian
Ordovician
Cambrian

PRECAMBRIAN

primitive gymnosperms, such as the ancestors of cycads,

Cretaceous

Permian

the uplands, but there were woodlands that included

Eocene
56

marine organisms, rich in calcium carbonate, became

Also during the Paleozoic, shifting seas created enor-

Neogene

Paleogene

were formed. Skeletal fragments from coral and other

mous swamps and tidal flats that were frequented by

0.01

Pleistocene

development of western landscapes.1 Unimaginable


during the Paleozoic Era (fig. 2.1), the area that would

AGE

65
145
200
251
299
318
359
416
444
488
542

Landscape History

from the older rocks and washed into the newly defined
inter mountain basins.
The Mesozoic would have been especially exciting
WYOMING

IA

S
LAURA

for biologists, as dinosaurs were abundant in the stillequator

tropical climate. About forty different kinds are recognized in the fossil record from Wyoming alone.5 Some
were huge, suggesting that the vegetation was abundant
enough to provide large amounts of food. Others were

GO

NDW
ANA

LAN

small, less than a few feet long. These reptiles, along


with other nondinosaurian reptiles like the flying pterosaurs, were important residents of Earth for more than

Fig. 2.2. Approximate location near the equator, about 240


million years ago, of the land area now known as Wyoming.
Adapted from Lageson and Spearing (1988).

100 million years. Fossils of some of the most famous


dinosaurs have been found at Como Bluff, between the
towns of Medicine Bow and Rock River. They include
Diplodocus, Apatosaurus, Stegosaurus, Allosaurus, and

were uplifted and essentially leveled by erosion within

Dryosaurus.6 Bison, pronghorn, and other familiar

tens of millions of years.

mammals would not be present for another 200 million

By the early Mesozoic Era, about 240 million years

years.

ago, most of Wyoming was slightly above sea level

Coexisting with the dinosaurs were flowering plants

and still near the equator (fig. 2.2). Fluctuating coastal

(angiosperms), such as magnolia, palm, fig, breadfruit,

environments in the Triassic Period led to formation of

sassafras, cinnamon, sweetgum, and willow trees, along

the colorful sandstones and shales of the Chugwater

with tree ferns, giant horsetails, clubmosses, and vari-

and Spearfish formations, both rich in iron oxides and

ous kinds of gymnosperms.7 Each of these types of

known as redbeds. Subsequently, in the Jurassic Period,

plants had various species, some adapted to the low-

extensive sand dunes were consolidated into the Nug-

lands and others to the uplands. Some coal deposits

get Sandstone. Various episodes of volcanism occurred

began to form during this time, coincident with the

toward the end of the Mesozoic, during the Cretaceous

dinosaurs, but most coal in the Powder River Basin did

Period, leaving deposits of fine ash across the landscape

not begin forming until the Paleocene, about 60 million

and in the sea.3 Much of this ash became bentonite, a

years ago, long after the large reptiles became extinct.

sodium-rich clay of considerable economic and ecological importance.

The Mesozoic ended with a cataclysmic event that


many believe led to the worldwide extinction of dino-

Also during the Cretaceous, thrust faulting created

saurs and numerous other forms of life. Known as the

the Hoback, Salt River, Sublette, and Wyoming moun-

Cretaceous-Paleocene Extinction Event,8 its cause is

tain ranges, now located near the Wyoming-Idaho

still debated, but there is little doubt that it was dra-

border. These mountains, known collectively as the

matic, possibly lasting only a few hours. A widely

Overthrust Belt and now an important source of oil

accepted explanation associates the extinction with a

and gas, are composed of sedimentary rock. Another

large asteroid impact on what is now the north shore of

round of mountain building began about 75 million

the Yucatan Peninsula. This occurred about 65 million

years ago, during the late Cretaceous, with folding

years ago, spewing huge quantities of dust into space,

and faulting of Earths crust throughout the region.

greatly reducing the amount of sunlight reaching Earth,

For about 30 million years, Precambrian granite and

which created cold temperatures that many species

other igneous and metamorphic rocks, together with

could not tolerate.9

overlying sedimentary strata, were thrust upward. This

But recently, an alternative explanation associates

period of mountain building is known worldwide as

the extinction with heat rather than cold.10 According

the Laramide orogeny. The overlying Paleozoic and

to this hypothesis, Earths surface was heated by the air-

Mesozoic sedimentary strata gradually were eroded

borne material resulting from the asteroids impact. The

13

14

Wyoming and the Rocky Mountain West

lofted dust and rocks fell to Earth through the atmo-

pine, redwood, spruce, sweet gum, sycamore, walnut,

sphere and were heated by friction to temperatures

and willow. Others, such as breadfruit and palm, are

higher than most unprotected animals could tolerate.

found today only in warmer climates. Some trees were

Fires would have burned across Earths surface wherever

large, up to 75 feet tall and 5 feet in diameter, as indi-

there was sufficient fuel. Supporting this hypothesis is

cated by fossil logs in Yellowstone National Park and the

the fossil record, which shows that, on land, only rela-

Shirley Basin south of Casper (see chapter 15).12

tively small animals survivedthose sheltered from

Coastal swamps were widely scattered during this

the heat in burrows or rock crevices. They included a

time, with cypress and other wetland treessimilar

variety of small mammals that had managed to coexist

perhaps to the cypress swamps of southeastern North

with the dinosaurs for millions of years. Also, protected

America today. Sphagnum bogs around the edges of lakes

by the water, aquatic life such as turtles and crocodiles

were also widespread. With a warm, subtropical environ-

survived. Could a comparable asteroid strike Earth

ment year-round, and an estimated annual precipitation

again? Over timespans of hundreds of millions of years,

of 100 inches of rain per year, plant growth was rapid.

the improbable becomes probable.

Moreover, the anaerobic conditions of the swamps and


bogs enabled preservation of dead plant material. With

The Age of Mammals


With the dinosaurs gone, mammals evolved to occupy

continued uplift and erosion from the adjacent mountains during the Laramide orogeny, the accumulating
peat was overtopped with deep sediments here and there

the vacated ecological niches. This marked the begin-

that compressed the plant material into coal. Today, most

ning of the Cenozoic. Early in this era, in the Paleocene

coal is too deep to mine, except where further uplifting

Epoch, the mammals included condylarths, creodonts,

and erosion occurred, leaving some of it near the sur-

tillodonts, pantodonts, and multituberculatesfor

face. That accounts for the abundant, easily accessible

which no common names can be found in the modern

coal in the Powder River Basinone of the worlds larg-

fauna. Most were small, described by paleontologists

est deposits. Some of the seams are more than 200 feet

as shrew-like, rodent-like, weasel-like, or beaver-sized.

thick. Because this coal developed largely in a freshwater/

Some of the herbivores had long, sharp front teeth,

estuarine environment, its sulfur content is low.13

possibly to assist with digging roots and tubers. They

Plant fossils in the Bighorn Basin have been studied

might have been omnivores. Though condylarths were

extensively. Scott Wing, a Smithsonian Institution paleo-

small in size, a branch of the group included the ances-

botanist, concluded in 1981 that 54 to 38 million years

tor of present-day hooved animals, such as horses, elk,

ago, during the Eocene, the basin had a subtropical cli-

and pronghorn. The pantodonts included the very first

mate with a dry season. The characteristic vegetation was

large mammals, which have been described as bear-

an evergreen broad-leaved forest with some conifers. For-

like or hippo-like herbivores. Early primates, rodents,

ests dominated by conifers most likely were characteristic

and coyote-sized carnivores also existed at this time in

of the mountains, but snowfall was probably a rare event,

the area that would become Wyoming. After another

and then only at high elevations. Many of the fossils are

10 million years of evolution, by the beginning of the

of plants that are common in the region today, whereas

Eocene Epoch, these forms, along with bats and oth-

others now occur naturally only in Southeast Asia, for

ers, had evolved into the verifiable early ancestors of

example, ginkgo and dawn redwood. Wing concluded

the mammals we recognize today.11

that the different kinds of plants were segregated into

Flowering plants also became increasingly common

different communities along gradients of temperature,

during the early Cenozoic. Some types of trees, shrubs,

water availability, and salinity. He estimated that the

and herbs found during this time, about 60 million

annual precipitation in the Bighorn Basin at that time

years ago, are still living in parts of North America and

was 60100 inches, much higher than the 424 inches

Eurasia. The trees included ancestors of alder, beech,

that falls there now. However, the shrub that became

birch, black locust, chestnut, cottonwood, cypress, dog-

known as greasewood was present, suggesting semi-arid,

wood, elm, fir, ginkgo, hickory, magnolia, maple, oak,

saline environments in some places.14

Landscape History

Even though the North American continent con-

Bighorn Mountains, Black Hills, and Uinta Mountains

tinued to drift poleward, the climate did not cool as

were formed. Another mountain range, known by geol-

would be expectedapparently because greenhouse

ogists as the Absaroka volcanic field, developed over a

gas concentrations in the atmosphere kept the region

shorter time, during the latter half of the Eocene. Dur-

about 18F warmer, on average, than today. Research

ing that time, the remains of entire forests became fos-

suggests that Earth warmed still further at the end

silized in what is now Yellowstone National Park, buried

of the Paleocene and beginning of the Eocene, about

by volcanic eruptions and subsequent flooding. Many

55million years ago. The warmest period, first reported

trees, now exposed by erosion, are still upright (see

in 1991 and now known as the Paleocene-Eocene

chapter 15).

Thermal Maximum (PETM), took about 10,000 years


to develop.

With erosion in the mountains and deposition in the

The PETM appears to have been associ-

adjacent basins, the topography was gradually flattened,

ated with an increase in the amount of carbon dioxide

broken here and there by hills of Precambrian granite

in the atmosphere, and it lasted for 100,000250,000

(fig. 2.3). Large rivers deposited prodigious amounts of

years. No mass extinctions of terrestrial organisms are

sediment during floods, and volcanism produced colos-

known to have occurred, though a large portion of lake

sal volumes of ash and volcanic debris. Both contributed

bottom protozoans disappeared. The climate became

to basin filling during the latter half of the Eocene and

significantly warmer and in some places more arid.

episodically throughout the Oligocene (encompassing

There would have been less plant growth. The fossil

45 to 23 million years ago). Also, huge freshwater lakes

record suggests that, with less food available, smaller

developed, covering 1025 percent of Wyoming at vari-

animals had a better chance of survival.

ous timesfirst Lake Lebo in the western Powder River

15

During the unusually warm PETM, the vegetation of

Basin (early Paleocene Epoch, 63 million years ago) and

present-day Wyoming shifted from dense forests, with

later Lake Tatman in the southern Bighorn Basin, Fossil

birch, sycamore, palm, magnolia, walnut, elm, laurel,

Lake near the eastern part of the Overthrust Belt, and

and dawn redwood, to more open dry forests, such as

Lake Gosiute in the Green River Basin and Red Desert

those found in parts of Central America today.16 Mem-

(early Eocene, 52 million years ago). Coal formed around

bers of the bean family, such as mimosa, became more

some of these lakes. Evidence for these lakes are fossilized

common, and the plants had smaller leavesan indica-

fish, now the primary attraction of Fossil Butte National

tion of a drier climate. Many tree species survived the

Monument near Kemmerer (fig. 2.4). Also found there are

PETM but were less common until after the heat sub-

fossils of stingrays, 13-foot-long crocodiles, banana trees,

sided, probably as carbon dioxide was reduced through

palm trees, water lilies, and numerous other plants and

storage in plant biomass, the oceans, and other parts of

animals.19 The climate was still warm and humid.

the biosphere.

Lake Gosiute is an example of how these now-extinct

The PETM is sometimes discussed in the context of

lakes affected todays vegetation patterns and land use.

current global warming, which is also associated with

This lakeshallow, saline, and without any mountains

increased concentrations of greenhouse gases. Much

nearbyfluctuated in depth as the climate changed

can be learned by studying the past, but Earth scien-

over 4 million years. It teemed with tropical life, in the

tists have calculated that the rate of warming today is

water and on shore. Eventually the Lake Gosiute eco

about fifteen times faster than at the end of the Paleo-

system led to the production of the worlds largest oil

cene.17 More rapid climate change provides less time for

shale and trona deposits, part of the Green River Forma-

organisms to adapt to their new environment through

tion. Trona is formed from sodium carbonate, sodium

evolution.

bicarbonate, water, and various impurities. As Craig

The Laramide orogeny, which began late in the Creta-

Thompson, professor at Western Wyoming College,

ceous, before the rise of mammals, continued well into

wrote, few passersby are aware that part of the steam

the Eocene. This mountain building produced modern

they see [from trona refineries near Green River] is water

configurations of the Wind River, Granite, Gros Ventre,

from Lake Gosiuteliberated and moving again in the

and Medicine Bow mountains (see fig. 1.2).18 Later, the

hydrologic cycle.20

15

Basin

Granite and
Related Rocks

Landscape History

As Lake Gosiute dried, continued erosion deposited


thick sediments over the salts and shales that had accumulated. The topography and climate continued to
change, until today the Chain Lakes wetland is located
in the bottom of the Great Divide Basin (see fig. 1.2),
surrounded by sagebrush steppe and desert shrubland
much different than the tropical life that existed there
in the past. The lakes occupy a small portion of the
depression where Lake Gosiute once existed.
Later, in the Miocene (about 10 million years ago),
regional uplifting became more pronounced, eventually bringing the landscape to near its present elevation
and initiating a new cycle of erosion. This uplifting was
accompanied by localized faulting and folding during
the Pliocene (5.0 to 2.6 million years ago) that led to the
formation of the Teton Rangethe youngest mountain
range in the region.21

Fig. 2.4. Fossil of a 16-inch fish (Diplomystus dentatus) that


lived about 50 million years ago in freshwater Fossil Lake, located near Kemmerer in present-day southwestern Wyoming.
Some fossil beds in this area are protected in Fossil Butte
National Monument. The fossils of palms and other tropical
plants and animals are found in the same area.

With uplifting, the climate cooled, and some rivers began to flow more rapidly, increasing the rate of

Pliocene, about 5 million years ago). 22 Drought-tolerant

erosion and eventually cutting canyons through the

plant species persisted or immigrated from nearby.

igneous rocks of buried mountain ranges (see fig. 2.3).

Some, such as sagebrush, needle-and-thread grass, and

Examples include the canyons of the Bighorn, Green,

bluegrass, migrated from the Old World across the Ber-

Laramie, Platte, Powder, Snake, Sweetwater, and Wind

ing Land Bridge that connected Siberia and Alaska at the

rivers. Most easily seen are the Snake River Canyon

time. Halophytic (salt-tolerant) plants found in modern

south of Jackson and the Wind River Canyon south of

inland salt marshes evolved from plants that are char-

Thermopolis. In addition to canyons, the accelerated

acteristic of marine coastal environments, such as salt-

erosion of claystones, siltstones, and other nonresistant

grass, alkaligrass, and cordgrass. Fires killed many of the

strata created badlands, a process that continues to this

trees that still grew on the upland, hastening the spread

day, carving escarpments in the elevated land.

of grassland and restricting woodlands to ravines, valley

As the mountains rose, the climate became more

bottoms, or ridges that burned less often. The grasslands

arid on the easternleewardside, caused by the rain-

created new ecological niches for mammals, and the pre-

shadow effect. This happens because the prevailing

decessors of all modern Wyoming animals appeared in

winds are from the west, and moisture is deposited on

the fossil record, including several kinds of pronghorn

the western slopes as the rising air cools and the mois-

that are known only from North America. Only one

ture condenses as it passes over the mountains. Along

species of pronghorn survives today. Also present were

with cooling, the drying climate caused the demise of

the camel, horse, mammoth, mastodon, rhinoceros,

some forests and favored the establishment or expan-

sabertooth cat, giant beaver, giant ground sloth, and

sion of shrublands and grasslands (beginning in the

short-faced bearall of which would become extinct in


North America about 10,000 years ago, a few thousand
years after the first humans arrived from Eurasia.

Fig. 2.3. (left) Intermountain basin development during the


past 70 million years. Ten to five million years ago the basins
were filled with sediments eroded from adjacent mountain
ranges. Since that time, rivers have cut canyons through
mountain ranges because of regional uplifting. Some basin
sediments have now eroded away, and the ancient mountain
ranges are again more fully exposed. See also chapter 17. From
Knight (1974, 1990). Drawings by Samuel H. Knight.

The Past 2 Million Years


Glaciers, warm interglacial periods, and erupting volcanoes characterized the Quaternary Periodthe past 2.6
million years. The first part of the Quaternary is known
as the Pleistocene Epoch or Ice Age. Snow accumulated in

17

Fig. 2.5.This landscape, with Gannett Peak in the background and Dinwoody Creek in the foreground, was buried
in ice about 21,000 to 16,000 years ago. Note the alpine
cirques and U-shaped valley, both formed by Dinwoody
Glacier. Woodlands of whitebark pine grow at treeline, and

forests of Engelmann spruce and subalpine fir occur on the


lower slopes. The riparian vegetation is dominated by various
species of willow (arctic, diamondleaf, Drummonds, grayleaf,
netleaf, and Tweedys ). At 13,809 feet, Gannett Peak is the
highest point in Wyoming. Photo by Scott Copeland.

Landscape History

Fig. 2.6 Jenny Lake in Grand Teton National Park is contained


by forested terminal and lateral glacial moraines at the base of
Cascade Canyon. The canyon is a U-shaped valley formed by
glaciers. The forests on the moraines are dominated by lodgepole pine, subalpine fir, and Engelmann spruce. Douglas-fir
and limber pine are common on the lower slopes of the

Tetons; subalpine fir and Engelmann spruce are abundant on


the north-facing slopes of the canyon, and whitebark pine is
common in the alpine zone. Shrublands dominated by mountain big sagebrush are found on the glacial outwash plains
below. The highest peak is Grand Teton (13,766 feet); the ele
vation of Jenny Lake is 6,779 feet. Photo by Michael Collier.

the mountains during cold periods, often forming drifts

about 7 times during the past 750,000 years (50 times

that persisted through the summer. More snow caused

since the mid-Pliocene). With each episode of melting,

higher reflectivity of solar radiation, lowering the tem-

widespread flooding shaped broad riparian floodplains,

perature further. Many snow masses were converted to

known as glacial outwash plains. Later these flood-

mountain glaciers.

plains became terraces and benches, such as along the

23

Far to the north, huge continental

glaciers moved southward, forcing the Missouri River

Snake River in Grand Teton National Park and along the

to flow into the Gulf of Mexico rather than into Hud-

Laramie River in the southeast (see chapter 4).

son Bay, as it had previously. Continental glaciers never

Huge freshwater lakes also formed, including Lake

reached Wyoming, but features such as terminal and

Bonneville in Utah, Lake Missoula in southwestern

lateral moraines, U-shaped mountain valleys, cirques,

Montana, and Lake Wamsutter in the Great Divide

glacial outwash plains, and kettle topography indicate

Basin of central Wyomingalong with the smaller Fre-

that extensive glaciers were present in most mountain

mont, Jackson, and Yellowstone lakes in northwestern

ranges and on the Yellowstone Plateau (figs. 2.5 and

Wyoming. 24 The grinding of rock by moving ice pro-

2.6). The nearest continental glacier was in present-day

duced fine glacial dust, which was washed down rivers

northern Montana.

and blown into the basins and plains. Deposition of this

During periods of glaciation, tremendous amounts

material, known as loess (pronounced lss), contributed

of water accumulated in the ice for thousands of years.

to the development of fertile soils in areas that now

Glaciers formed and then thawed during warm periods

support grasslands and shrublands. As the snow melted

19

20

Wyoming and the Rocky Mountain West

each spring, mountain soils were subjected to flushing

Shoshoni, Sioux, and Ute.31 By that time the glaciers had

by large volumes of water, which eroded soil particles

receded into the mountains.

and slowed the accumulation of nutrients. Soil develop-

Learning about change during the Holocenethe

ment in the mountains was most rapid where soft rocks

past 10,000 yearshas been facilitated by the analysis

(such as shales) were exposed instead of harder rocks,

of fossil pollen and other plant fragments. In wetlands,

like granite, rhyolite, limestone, and some sandstones.

these materials settled and were preserved in anaerobic

Volcanic activity during this time included enormous

sediments for thousands of years. 32 Paleoecologists

caldera-forming eruptions followed by multiple lava

have learned to identify the kinds of plants from which

flows that created much of the topography seen in

the partially preserved pollen, cones, seeds, and leaves

Yellowstone National Park today (see chapter 15).25

came. Because sediments accumulated with time, the

Though commonly thought of as the Ice Age, the

deeper plant materials usually are older. Thus, changes

Pleistocene also had extended warm periods, even

in plant materials with depth suggest how the vegeta-

warmer than during much of the past 10,000 years. 26

tion of the surrounding land area has changed over

All or most of the ice melted at such times. Present-day

thousands of years.

glaciers formed about 400 years ago, during a period

Paleoecological studies in Wyoming and nearby pro-

known as the Little Ice Age (ca. 15501850). They are

vide evidence of continued change through the Pleis-

now small, confined to high mountain peaks, espe-

tocene to the present (fig. 2.7). 33 For example, plant

cially in the Wind River Mountains, and are now reced-

fragments collected from the northern end of the Teton

ing due to climate warming during recent decades (see

Range document the presence of ponderosa pine about

chapter 3).27

127,000 years ago, during the last interglacial period,

At the beginning of the Quaternary, 2.6 million

when the climate was warmer than it is today. This coni-

years ago, the flora of the region was similar to what it

fer is now absent from that area and the Yellowstone

is today.28 Forests dominated by familiar conifers were

region. Studies by Cathy Whitlock and her students at

common in the mountains, and shrublands and grass-

Montana State University describe the vegetation his-

lands were common in the basins and on the plains.

tory of this area over the past 30,000 years, during and

Cottonwood, aspen, bur oak, boxelder, paper birch, and

following the last glacial period (known as the Pine-

Rocky Mountain maple were present, but most other

dale). They concluded that ice covered the Yellowstone

species of broad-leaved trees, along with redwood and

region at times during the Pleistocene, 34 when the mean

cypress, have not been found in the fossil record for that

annual temperature was about 918F lower than it is

time.

today. Conifers were restricted to a narrow band in the


foothills, and tundra-like vegetation was widespread

Thousands of Years Ago

across the lowlands (fig. 2.8). Throughout Wyomings


basins, permafrost was widespread. Several of the preva-

The first people arrived in North America about 15,000

lent mammals at the time are now characteristic of the

years ago.

With their ingenuity, tools, and weapons,

Arcticcaribou, collared lemming, and barren-ground

they spread rapidly. Their hunting is commonly thought

muskox.35 Mammoth and mastodon roamed the inter-

to have been a primary cause of the extinction of large

mountain basins as well.

29

mammals, such as the mammoth, giant sloth, camel,

Whitlocks

research

indicates

that

Engelmann

horse, and various other species that existed at the time;

spruce was one of the first trees to colonize the areas

but contributing factors may have been climate warming

vacated by receding glaciers, but within a few centu-

and reduced food availability caused by droughts. The

ries it was joined by whitebark pine, subalpine fir, and

hunters became known as the Clovis people. They were

lodgepole pine. Such forests were widespread at middle

able to kill 10-ton mammoths with their stone-tipped

and high elevations throughout Wyoming until 11,000

weapons.30 Clovis descendants eventually organized into

years ago, when the climate became warmer and drier.

various tribes, including the Apache, Arapahoe, Black-

These conditions led to an upward shift in upper and

foot, Cheyenne, Comanche, Crow, Flathead, Kiowa,

lower treelines and the development of lodgepole pine

Landscape History

PINE
SAGEBRUSH
wetter

14

C yr
(ybp)
B.P.

3,520

SAGEBRUSH
PINE
drier

8 7 6 5 4 3 2 1 2 3 4 5 6 7 8 9 10 11 12

5,850

HOLOCENE

8,460
10,060
10,200

TRANSITION

25,540
22,300
20,880

21,890

PLEISTOCENE

21,190

26,640
Fig. 2.8. Fossil ice wedge exposed at the Rawlins landfill,
which suggests that a tundra-like environment existed during
the Pleistocene (Mears 1981, 1987). The rod is 3 feet long.
Elevation 6,900 feet. Photo courtesy of Brainerd Mears, Jr.

>29,000
TEPHRA
61,000 ybp

and Douglas-fir forest at middle elevations. Forest fires


became more common. In the past 6,000 years, the climate has generally cooled, allowing spruce, subalpine
fir, and whitebark pine to expand their range. This trend
was reversed during brief, warm, dry periods, such as
the Roman and Medieval warm periods, about 2,000
and 1,000 years ago, respectively. These intervals of past

Fig. 2.7.The ratio of pine and sagebrush pollen has changed


during the past 60,000 years. A higher peak on the left indicates that pine pollen was more abundant than sagebrush, an
indication that the climate was wetter; a higher peak on the
right indicates that sagebrush pollen was more abundant. Based
on data from Grays Lake, southeastern Idaho. 14C yr B.P. = years
before present as determined by carbon-14 dating. Adapted
from Beiswenger (1991).

drought provide important insights on how ecosystems


will respond to the warmer, drier conditions projected
for the future (see chapter 3).
A comparatively warm, dry period spanning 3,000
years in the early Holocene caused an expansion of
big sagebrush, greasewood, juniper, and grasses in
the lowlands. During this time, grasses covered some
north-facing slopes up to 7,900 feet elevation in the
mountains.36 Most mountain tree species, such as
Engelmann spruce, subalpine fir, and lodgepole pine,
survived at higher elevations, though they probably

21

22

Wyoming and the Rocky Mountain West

also existed in relatively cool, moist valley bottoms at

the past 2 million years or so, now it is clear that Utah

lower elevations. Droughts during this time killed some

juniper arrived in southwestern Wyoming about 9,000

of the grasses that stabilized sand dunes in the region,

years ago, having migrated roughly 200 miles north-

resulting in more blowing sand and shifting dunes.

ward as the climate warmed following glaciation. 39 Even

There is evidence that the Platte and other large rivers

more surprising, ponderosa pine did not reach the Big-

were sometimes dry during this period. 37

horn Mountains of Wyoming until about 3,000 years

The work of paleoecologists also has been facilitated

ago.40 Pinyon pine did not arrive in the state until about

by the middens of bushy-tailed woodrats, often referred

700 years ago and is still found naturally only where

to as packrats, because they collect leaves, seeds, cones,

the Green River flows into Utah.41 This small tree also

bones, and other material and deposit them in their

occurs in northern Colorado, just south of Owl Canyon

nests (fig. 2.9). Not particularly tidy, the woodrats uri-

on the east side of U.S. highway 287.

nate and defecate on their middens time and time again,

University of Wyoming paleoecologist Stephen T.

creating hard masses that often persist for thousands of

Jackson and his colleagues studied middens in the

years. By identifying the plants represented in differ-

foothills of the Pryor Mountains, at an elevation of

ent layers of a midden, and studying middens of dif-

about 5,000 feet on the north side of the Bighorn Basin

ferent ages, scientists draw conclusions about how the

(fig.2.10).42 They observed the following changes. About

vegetation has changed where the woodrats have for-

10,000 years ago, when glaciers were retreating up the

aged. Radiocarbon dating provides a rather precise time

mountains, the climate in the foothills was cool and

frame, with one midden in northeastern Utah having

humid, as indicated by the presence of ground juni-

deposits that are 40,000 years old. 38 Studying vegeta-

per, Rocky Mountain juniper, chokecherry, wood rose,

tion change using pollen or middens found at treelines

and fernsall of which today are found primarily in

is especially instructive, because trees are sensitive to

relatively moist woodlands and shrublands. In more

water stress. Thus, climate change is easily detected as

shallow layers of the middens, deposited about 8,000

the tree fragments appear and disappear from top to

years ago, fragments of drought-tolerant plants were

bottom in the middens.

foundnamely, pricklypear cactus, saltbush, winterfat,

Middens have provided some surprising informa-

and later, Utah juniper. The plants represented in older

tion. First, although it is commonly thought that most

midden deposits were absent. From about 4,400 to 2,700

of the same plant species have been in Wyoming for

years ago, limber pine expanded into the area, for reasons that are not yet clear.43 About 2,700 years ago, more
arid conditions returned, as indicated by the absence of

Fig. 2.9. Bushy-tailed woodrats are 1118 inches long. About


half their length is tail. The rodents gather identifiable plant
materials for their nests, which may persist for more than
10,000 years. The nests are known as middens. Photo by Phil
Dotson/Science Source.

limber pine, presumably because it could not survive the


droughts. Utah juniper and curlleaf mountain-mahogany became more common, both of which are abundant
around the middens now.
This Bighorn Basin research hinges to a large degree
on the presence or absence of the various species of
shrubs and trees at lower elevations. Another study,
conducted in the Wind River Mountains by Patricia Fall
and her associates from Arizona State University, studied changes in alpine treeline using pollen data from
a lake at about 10,200 feet.44 They found that, prior to
about 11,300 years ago, with glaciers still nearby, the
climate was relatively cold, and the vegetation around
the lake was largely alpine tundra. However, by about
10,600 years before present, forests of whitebark pine,
Engelmann spruce, and subalpine fir surrounded the

Landscape History
Big Pryor

East Pryor

is often mentioned during discussions about causes of

U
ta
h

R
oc
ky

tn
.J
un
ip
Ju
er
ni
p
e
Ju
r
ni
pe
rp
ol
le
n
R
oc
ky
M
tn
.J
U
ta
un
h
ip
Ju
er
ni
p
e
Ju
r
ni
pe
rp
ol
le
n

present-day global warming (see chapters 3 and 18).46

landscape change by studying tree rings, soil profiles,

2000

anthropological data, and old journals and photographs. What do such records reveal about the past few

3000

Age (14C yr B.P.)

In addition to the study of pollen and plant fragments,


scientists from various disciplines have learned about

1000

centuries?

4000

First, tree-ring studies have documented ongoing


climate change (fig. 2.11), including droughts during

5000

8001300

6000

ad

that were far more severe than any since

that time. Tree-ring data collected by Steve Gray and his


associates at the University of Wyoming indicate that

7000

these extended droughts lasted 75 years or more and

8000

were long enough to cause changes in soil profiles.47 Soil


characteristics suggesting drought are associated with

9000
10000

Centuries and Decades Ago

fewer American Indian artifacts, perhaps because the


5

80

80

Relative abundance (percent)


Macrofossils (solid bars), pollen (open bars)
Fig. 2.10. Abundance of plant fragments from two species of
juniper at different times during the past 10,000 years, based
on data from woodrat middens at two foothill locations on
the north side of the Bighorn Basin. The length of the solid
horizontal bars indicates the proportion of all plant fragments
that originated from a single species, which reflects the abundance of each species in the surrounding area. The open bars
indicate the relative abundance (percent) of juniper pollen
(not distinguishable by species). 14C yr B.P. = years before present as determined by carbon-14 dating. Adapted from Lyford
et al. (2002).

number of bison was lower at such times. Since about


1300

ad,

Wyoming has experienced a relatively moist,

cool climate. Paleoecologists refer to the 300-year period


from about 1550 to 1850 as the Little Ice Agea period
when a 2F cooling of mean annual temperature was
apparently enough to cause further vegetation and soil
changes at the boundary between grasslands and big
sagebrushdominated shrublands.48 Grasslands were
more widespread during relatively warm, dry periods,
because sagebrush is less drought tolerant (see chapter7).
Compared to changes during the eras and epochs
of the more distant past, Wyoming landscapes have
changed relatively little during the past few centuries.
American Indians surely influenced the landscape,

lake. Moreover, the treeline was probably 300 feet

but the details are not well documented for most of

higher than it is today, suggesting that the climate was

the area.49 We can infer that they used small trees for

warmer at that time than it is today. This may have

building lodges, and that they started fires from time

persisted until about 3,000 years ago. Since then, the

to timesome burning over large areas. The adoption

alpine lake has been cooler and surrounded by a mosaic

of the horse in the 1700s greatly increased hunting

of subalpine meadow and tundra with scattered white-

efficiency. That and human-ignited fires probably had

bark pine, as is typical of the transition from subalpine

the most pronounced effect on western ecosystems.50

forest to tundra in this region. Dense forests persisted

Also, some 65,000 emigrants passed through Wyoming

only at mid-elevations, as they do today. Lodgepole

in 1850 alone, some bound for California to search for

pine became abundant during the past 3,000 years.45

gold, others headed to Oregon Territory for land and

Such studies indicate that the climate has changed

other opportunities.51 Military forts were built along

dramatically during the past 10,000 years, a fact that

the way. Some travelers decided to live in Wyoming

23

Wyoming and the Rocky Mountain West

Proportion of western
U.S. in drought (%)

24

0
10
20
30

Wetter

40
50
60

Little Ice Age

Drier

70
800

1000

1200

1400

Year AD

1600

1800

2000

Black areas indicate drought. Note that the past 400 years
have been relatively moist. Long droughts of at least several
decades occurred before 1300 ad and were more severe than
the droughts of, for example, the 1930s. See also Wise (2012)
and fig. 3.5. Adapted from Gray et al. (2006).

Fig. 2.11. Drought occurrence and duration in the interior


western states during the past 1,200 years based on tree-ring
data collected from sites throughout the region. The horizontal line indicates the average proportion of the western
United States that experienced drought during this time.

Territory, becoming ranchers, farmers, miners, mer-

outstanding example and was used actively by Indi-

chants, railroad workers, telegraphers, educators, and

ans from about 500 to 200 years ago.55 Hunting pres-

industrialists.52

sure by Indians intensified after they acquired horses.

The first significant influence of EuroAmericans was

Even more bison were killed when the railroad arrived

53

in 18671868, with the tendency of early travelers to

Such trappers as Wilson Price Hunt, Jedediah Smith,

shoot large numbers of animalspart of the adven-

and Jim Bridger were lured to the region because beaver

ture. 56 Bison were essentially absent as a free-roaming

pelts were highly valued for the fabrication of felt used

animal by the 1880s, about the same time that domestic

through the trapping of beaver in the early 1800s.

for clothing, primarily mens hats. Their unregulated

livestock herds were increasing. By 1900 the population

trapping caused a sharp decline in beaver popula-

of all large mammals (bison, mule deer, whitetail deer,

tions. Soon, many beaver dams failedcausing soil

elk, pronghorn antelope, bighorn sheep, and black and

erosion and the development of gullies in some areas

grizzly bear combined) in Wyoming had been reduced

where ponds, meadows, and willow-dominated shrub-

to as few as 60,000. Only with the conservation move-

lands had existed previously (see chapter 4). Some early

ment fostered by President Theodore Roosevelt and oth-

explorers noted how the gullies could be a madden-

ers in the early 1900s did natural resource management

ing hindrance that would require traveling a circu-

improve so that big game populations in Wyoming

itous route of several miles just to make one mile in the

could recover. In 2013, the Wyoming Game and Fish

desired directioneven though the gullies were only

Department estimated that the state was providing hab-

a few yards wide. Dudley Gardner, author and arche-

itat for 528,000 pronghorn, 427,000 mule deer, 104,000

ologist, suggests that most gullies were the effect of

elk, 60,000 whitetail deer, 7,400 moose, 5,500 bighorn

trapping too many beaver, though he acknowledged

sheep, and about 4,000 bison (including Yellowstone

that other factors could have been involved, such as

National Park). For comparison, the state had about

an increase in wagon traffic and overgrazing by horses

576,000 people. 57

and cattle.54 Notably, the market for beaver pelts dimin-

A useful approach for determining landscape changes

ished rapidly in the late 1830s. Beaver are now common,

during the past 100200 years is to compare present-

though much less so than previously.

day conditions with those described in early journals.

Indians had subjected bison and other animals to

Robert Dorn, noted Wyoming botanist, ornithologist,

hunting for thousands of years, often using cliffs to

and natural historian, summarized salient observations

kill or trap their prey. The Vore Buffalo Jump, located

by trappers, explorers, and settlers in Wyoming during

along Interstate 90 in northeastern Wyoming, is an

18051878 and concluded the following:58

Landscape History

1. Bison were present in all parts of Wyoming, includ-

mer of 1886 and the devastating winter of 18861887,

ing subalpine meadows, but they were especially

which killed 4060 percent of the cattle in most of

common on the eastern plains. Evidence of heavy

Wyoming.62 Vegetation changes caused by livestock in

grazing by bison was common.

the late 1800s and early 1900s are difficult to assess,

2. Range fires occurred, but they were not mentioned


often; grazing by native animals may have mini-

but they surely were substantial and are still apparent


in some areas (see chapters 6 and 7).

mized fuel accumulation.

Forests also were subjected to new kinds of human

3. Forest fires were common.

influences in the 1800s. The Union Pacific Railroad, the

4. Big sagebrush was very common.

nations first transcontinental railroad, required rail-

5. Grizzly bears, black bears, wolves, and elk were com-

road ties, and lumber and timbers were needed along

mon in grasslands and shrublands of the basins and

the way for new towns and for mining gold, silver, and

plains but declined with increased hunting pressure.

copper. The remnants of slab piles can still be found,

6. Some streams were ephemeral and had steep-walled

indicating the locations of sawmills. Some forests were

gullies.
7. Grasshopper and Mormon cricket plagues occurred,
as they do now.
8. Vegetation patterns have changed very little since
the mid-1800s.
Human influences were minor when the early explorers were writing their journals in the 1800s. 59 Probably
fewer than 10,000 people resided in Wyoming Territory
at the time, less than 10 per square mile. However, more
than 350,000 immigrants traveled through Wyoming
along the Oregon Trail between 1841 and 1868, and
with the wagon trains came thousands of livestock.
These animals grazed along the way and sometimes were
concentrated in small areas. For example, one group of
cowboys worked unsuccessfully for five days to drive its
herd across the Platte River near Casper, and during this
time the grass was devoured for miles around.60 During the 1800s, the environmental impacts of people and
livestock were probably much greater along the Oregon
Trail than in the region as a whole.

essentially clearcut. Elsewhere, forests were thinned as


the tie-hackers cut only the trees they wanted. Railroad
ties typically were floated down creeks and rivers to railheads, especially during spring when water levels were
high. Such activity, along with placer mining for gold,
must have caused significant changes along some rivers
and creeks.
The first photographers provided another kind of historical record. Several studies have compared early and
modern photographs of the same areas. For example,
rangeland ecologist Kendall Johnson retraced the route
of William Henry Jackson, the famous landscape photographer who accompanied the 1870 expedition of
geologist Ferdinand Hayden across Wyoming, which
occurred prior to the cattle boom.63 Johnson located
and rephotographed 56 scenes in the 1980s, approximately one century later. The photo comparisons led
him to conclude:
1. Except in the relatively small area where the land had
been plowed, the grasslands of eastern Wyoming are

The effect of livestock became more widespread

very similar to those of 100 years ago. Overall, the

after the completion of the Union Pacific Railroad in

effects of livestock grazing appear not to have been

1869, which made western rangelands a source for

much different than the effects of native ungulates,

much of the nations beef, lamb, and wool. Large herds

although shifts in plant species composition may

of cattle were driven northward from Texas. Sheep were

have occurred.

imported by railroad, and some herds grew to 10,000

2. Big sagebrush has increased, decreased, or remained

animals. Investors and western stock growers became

about the same, depending on land use and site char-

wealthy. Livestock were forced to graze on marginal

acteristics, but it was abundant in the intermountain

rangelands, leading rangeland scientists John Mitchell

basins long before livestock arrived.

and Richard Hart to write, the land was filled entirely

3. Woodlands and forests on the uplands have become

with cattle.61 Fortunately for rangeland condition,

more dense, probably because of fire suppression,

the livestock boom ended in 1887, following a crash

but trees have not invaded adjacent vegetation types

in the beef market that coincided with the dry sum-

over large areas.

25

26

Wyoming and the Rocky Mountain West

4. River bottoms have changed dramatically, caused

agricultural purposes, often accelerating erosion. Live-

by impoundments, irrigation, cultivation, livestock

stock have replaced bison almost everywhere, and many

grazing, and settlements; and some river margins

plant species have been introduced from other conti-

now have more trees than in 1870.64

nents, some becoming invasive weeds. Towns, cities,

5. Areas of sparse plant growth, such as desert shrub-

highways, coalmines, oil and gas fields, refineries, wind

lands, retain a nearly identical appearance to those

energy projects, power lines, dams, and reservoirs have

of 1870.

been built. Riparian landscapes and some grasslands

Photographs must be used cautiously in drawing conclusions, as they only provide snapshots in time of a

and forests have been altered most dramatically; trees


have been planted where there were none before. Some

small area.65

industrial impacts have accelerated to alarming rates in

Since Wyoming became a state in 1890, changes have

and western Great Plains. Still, many parts of Wyoming

continued to occur. Forest landscapes in many areas

and nearby states are little changed since American

have been fragmented by road building and timber

Indians first encountered white-skinned explorers from

harvesting, and some rangelands have been plowed for

the east.

the past 20 years, especially in the intermountain basins

Present-Day
Environments and Climate
Change
Chapter 3

Describing the environment in a way that is meaningful


for understanding the growth and abundance of plants,
animals, and other forms of life is one of the greatest
challenges faced by ecologists. Convenient numbers
are sometimes calculated, such as monthly or seasonal
averages of temperature, but organisms respond to the
full complex of interacting environmental factors
not to one individual factor or to statistical summaries. Moreover, extremes are usually more important
than averages, and young individuals are less tolerant
of extremes. Therefore, the distribution of a species
depends more on the environment prevailing when the
plants or animals were youngconditions that usually
are not known when the focus is on adult plants. Little
wonder, then, that correlations between environmental
variables and the presence or absence of a species are
sometimes hard to find. The challenge is to integrate
the diverse factors that determine whether an organism
will survive.
The environments of wetlands, grasslands, shrublands, forests, tundra, and other kinds of ecosystems are
discussed in subsequent chapters. Here, general patterns
are described, with emphasis on the effects of topography and elevation on precipitation, temperature, evaporation, growing season length, soil characteristics, and
the nature and frequency of periodic disturbances. All
of these variables interact in determining the nature of
the ecosystem in a specific areaand all are affected by
climate change, the topic of the last section.

Topography
Wyoming landscapes span an elevational range of 10,758
feet, from the lowest point on the plains at 3,125 feet,
where the Belle Fourche River flows into South Dakota,
to the top of Gannett Peak at 13,809 feet. More than
one-third of the state is above 7,000 feet, and 10 percent is above 9,000 feet (fig. 3.1). About fifty mountain
peaks are 13,000 feet or above, mostly in the Wind River
Mountains. Several major rivers originate in the state
and flow into the Missouri, Colorado, or Columbia river
drainages (see chapter 4).
Elevation affects all environmental variables. Temperature decreases as one climbs, whereas annual precipitation generally increases.1 These trends create a
cooler and relatively moist environment in the mountains. Some groups of closely related plants have different species adapted to lowland, foothill, montane, and
alpine environments (fig. 3.2). The effects of elevation,
however, are strongly moderated by the influences of
topography and the increased rates of evaporation that
occur at higher elevations because of lower atmospheric
pressure in the mountains.2 Thus, high south-facing
slopes that receive direct solar radiation throughout the
year sometimes are as dry as low-elevation deserts. In
contrast, north slopes at low elevationscooler because
of less direct sunlighthave vegetation normally found
at higher elevations. Such observations have led ecologists to conclude that topographic position and degree
of exposure to direct sunlight are more important than

27

28

Wyoming and the Rocky Mountain West

Sheridan

Cody

Powder River Basin

Big Horn
Basin

Gillette

Worland
Newcastle

Wind River
Basin

Casper

Lander

Great
Divide
Basin

Green
River
Basin

Rawlins

Rock
Springs

Laramie
Cheyenne

9,000 ft

7,000 ft

5,000 ft

Fig. 3.1. General topographic map of Wyoming with the


names of selected towns and intermountain basins. Elevation
is given as number of feet above sea level. See fig. 1.2 for the

elevation in determining plant and animal distribution


patterns.

0
0

50
50

100 km
100 miles

names of mountain ranges and other basins. Adapted from


Martner (1986).

To the west of the Great Plains is a region of mountains and intermountain basins (see figs. 1.2 and 3.1).

Based on elevation, topography, geologic charac-

The Continental Divide passes through the middle

teristics, and location with regard to the Continental

of this region, splitting west of Rawlins to form the

Divide, several ecological regions can be identified. The

Great Divide Basin, which has no drainage to either

Great Plains characterize the eastern third of Wyoming,

the Pacific or Atlantic oceans. The Divide around this

with elevations ranging from 3,125 to 7,524 feet. Sum-

closed basin is relatively low, about 7,000 feet above sea

mer precipitation is more frequent on the plains than

level, and is characterized by sagebrush steppe rather

farther west. Grasslands are the most common vegeta-

than the coniferous forest or alpine tundra typical to

tion type, and there is evidence to suggest that bison

the north and south. For this reason, the Pony Express,

were more numerous there than in the Great Basin of

various wagon trails, and the first trans


continental

Utah and Nevada. With more bison, plants on the

railroad passed through the area in the 1800s. Most of

Great Plains evolved adaptations to withstand grazing

the basins range in elevation from about 3,960 to 7,260

from large herbivores; hence, they are more tolerant of

feet, with the highest basins in the south (see figs. 1.1

livestock grazing than are plants in the Great Basin.

and 3.1).

Present-Day Environments and Climate Change

ELEVATION (m)
1000

2000

3000
Fig. 3.2. Different species of the same plant
genus are often found at higher or lower
elevations, indicating different adaptations to
the varying environmental conditions along
mountain slopes. Note that some species have
broader environmental tolerances than others.

Pinus spp.
Whitebark pine
Lodgepole pine
Ponderosa pine
Pinyon pine
Limber pine
Artemisia spp.
Alpine sagewort
Mountain big sagebrush
Wyoming big sagebrush
Black sagebrush
Basin big sagebrush
Silver sagebrush
Salix spp.
Arctic willow
Planeleaf willow
Pacific willow
Coyote willow
Peachleaf willow
Elymus spp.
Scribner wildrye
Blue wildrye
Canada wildrye
Stipa spp.
Columbia needlegrass
Needle-and-thread
Green needlegrass
Poa spp.
Alpine bluegrass
Big bluegrass
Canby bluegrass
Sandberg bluegrass
3000

5000

7000

9000

11000

ELEVATION (feet)

Scattered throughout the Great Plains and inter-

Some of the more erodible sedimentary bedrocks,

mountain basins, and in the foothills of the mountains,

such as shales and siltstones, lead to the formation of

are topographic features best referred to as escarpments

classic badland topography (for example, Adobe Town

or ridges. They are usually composed of erosion-resistant

southeast of Rock Springs, Gooseberry Creek badlands

sandstones and limestones (fig. 3.3). The vegetation is

east of Meteetsee, Hells Half Acre west of Casper, the

distinctive, primarily because of a coarse, rocky sub-

Honeycomb Buttes northeast of Rock Springs, and the

strate with very little soil and because of drifting snow.

Wind River badlands near Dubois). Other topographic

Shrublands dominated by mountain-mahogany, big

features in the lowlands are sand dunes, arroyos (gul-

sagebrush, and skunkbush are common, as are wood-

lies), buttes, playas, deflation hollows, mima mounds,

lands characterized by juniper, limber pine, and pon-

glacial moraines, hogback ridges, bajadas (alluvial and

derosa pine (see fig. 1.5).

colluvial fans), canyons, nivation hollows, and braided

29

30

Wyoming and the Rocky Mountain West

Fig. 3.3.On this sandstone escarpment, limber pine and


Rocky Mountain juniper grow only in cracks along fracture
planes where soil, water, and nutrients are available. This escarpment is part of the Frontier Formation and is located near

Muddy Gap, north of Rawlins and west of the Ferris Mountains. Wyoming big sagebrush and greasewood shrublands
are found on the adjacent fine-textured soils. Photo by Ken
Driese.

or meandering stream channels. The ecological sig-

sure of water availability.4 Where precipitation is higher,

nificance of such features is a recurring theme in this

such as in mountain forests, other factors may be more

book.

limiting (for example, temperature or the length of

Constituting another third of Wyoming, the moun-

the growing season). In all cases, water availability to

tains are characterized by lakes, streams, coniferous

plants is affected as much by the infiltration and water-

forest, aspen groves, subalpine meadows, and alpine

holding characteristics of the soil as by the amount and

tundra. Environmental factors vary greatly, with freez-

timing of precipitation.

ing temperatures possible every month of the year at

Throughout Wyoming and much of the region, the

high elevations. Geographers have classified the Lara-

mean annual precipitation varies by a factor of 10, from

mie, Medicine Bow, and Sierra Madre mountains in the

6 to 60 inches (about 15150 cm; fig. 3.4). The two dri-

southeast as part of the Southern Rocky Mountains and

est areas are the Great Divide Basin, near Wamsutter

most of the other mountain ranges as part of the Middle

on Interstate 80, and the northern part of the Bighorn

Rocky Mountains (see fig. 1.2). Only the Beartooth,

Basin, between Lovell and Powell. In general, the inter-

Madison, and Gallatin ranges of the Greater Yellow-

mountain basins in the western two-thirds of the state

stone Area are in the Northern Rockies.

are drier than the Great Plains to the east, with averages
of 612 and 1216 inches per year, respectively. The

Precipitation
As in other semi-arid environments, precipitation has

foothills and mountains receive 1660 inches, with the


highest annual precipitation in the southwestern corner
of Yellowstone National Park.

a significant effect on plants and animals, as well as

With prevailing westerly winds, the western slopes

on human activities. In a given temperature zone, the

of the mountains receive relatively more precipita-

amount of plant growth in grasslands and shrublands

tion, the result of condensation as air moves upward,

is typically correlated with rainfall or some other mea-

cooling as it crosses the mountains. The leeward

Present-Day Environments and Climate Change


Fig. 3.4. Average annual precipitation, 19812010. From
Wyoming Water Resources
Data System, prepared with
PRISM Climate Group data,
Oregon State University
(http://www.prism.oregonstate.edu/terms.phtml).
Cartography by Christopher
Nicholson.

slopes and basins are drier than might be expected,

mountains varies greatly from one year to the next.

because descending air masses become warmer, which

This variation in Wyoming is not well correlated with

increases the potential for evaporation instead of pre-

El Nio and La Nia years, as seems to be true farther

cipitation. This is the well-known rainshadow effect.

to the south and north.6

Most Wyoming basins and plains are in a rainshadow,

The form, duration, and intensity of precipitation

with the relatively dry conditions ameliorated during

are as important as the amount. Rain that occurs dur-

years when more than average amounts of moist air

ing warm periods evaporates quickly, sometimes even

from the Gulf of Mexico reach the plains and moun-

before it wets the rooting zone of the soil. Heavier

tains. Similarly, a northerly flow of moist air from

rainstorms can cause infiltration down to 4 inches

the Gulf of California (the Arizona Monsoon) causes

or more, but plant water stress soon develops if sig-

added summer precipitation in southern Wyoming, a

nificant rainfall events are separated by more than

phenomenon that could account for the presence of

a week or two, as they usually are. In contrast, snow

Gambel oak and the extensive groves of aspen on the

common
ly accumulates during winter. Some snow

west slopes of the Sierra Madre (see chapter 10), where

will be sublimated, but there is the potential for

they grow much as they do in parts of Colorado and

deeper infiltration into the soil when the snow melts,

Utah. Some years have high amounts of precipitation,

at a time when less evaporation occurs because of

generally caused by flows of such air masses from the

lower temperature. Such deep percolation and storage

south and southeast. Other years are extremely dry,

enable plant growth well into the summer and seem

occurring when these air masses are diverted to the

important for the survival of big sagebrush in most

east. As with rain in the lowlands, snowfall in the

intermountain basins. Not surprisingly, big sagebrush

31

Wyoming and the Rocky Mountain West

Palmer Drought Index

32

wet
Wet

Little Brooklyn Lake

Early 20th Century:


Wetness

Pre-1900: Multiple droughts >15 years


many of greater severity than 1950s

10361 ft

48 in

1
0
1
2
3

Dry
dry
900

Late 20th Century:


Relatively wet &
no major droughts

50+ year
megadroughts
1000

1100

1200

1300

1400

Year

1500

1600

1700

1800

1900

2000

Foxpark
9065 ft

Fig. 3.5. (above) Drought history for the past 1,100 years in the Green River Basin of southwest
Wyoming, as reconstructed from tree rings. The graph shows estimated values for the Palmer
Drought Severity Index, a measure of drought severity. Positive values of the index represent
relatively wet conditions; negative values indicate drought. Each point on the graph represents the mean over a 25-year period. See also fig. 2.11. Based on data from Cook et al. (2004);
adapted from Gray and Andersen (2009).
Fig. 3.6. (right) Climate diagrams for southeastern Wyoming that illustrate climate change with
elevation. The graphs show mean monthly precipitation (top line) and mean monthly temperatures from January to December. Black areas indicate periods of drought, when the precipitation
line drops below the temperature line. Each unit on the vertical axis is 20 mm of precipitation
(water equivalent) or 10C. The number at the top of each diagram on the left is elevation (feet);
on the right, mean annual precipitation (inches). A break in the horizontal bar in the summer
indicates the frost-free period; a black bar indicates that the mean daily minimum during a
month is below freezing, and a hatched bar indicates months when the lowest temperature is
below freezing. On the vertical axes, to convert millimeters of precipitation to inches and degrees Centigrade to degrees Fahrenheit, see appendix A. Based on data from Martner (1986) and
the Wyoming Water Resources Data System at the University of Wyoming.

Laramie
7267 ft

Cheyenne
6125 ft

occurs most abundantly in the basins of the western


two-thirds of the state, where the proportion of snow
is higher.7 Proportionately less snow and more spring

as the lowlands, with those of the 1930s, 1950s, and


the first decade of this century lasting up to 10 years.8
Much longer droughts have been identified using treering research in Wyomings Bighorn Basin, with some
lasting 50 years or more prior to 1700 (fig. 3.6).9 Could
multidecadal droughts occur again? Understanding the
effects of drought on ecosystems and local economies
is a logical step toward developing a strategy for adapting to the droughts of the future. Even today, water is a
limiting factor for economic growth.

13 in

C
0

(fig. 3.5), where big sagebrush often is restricted to


Periodic droughts characterize the mountains as well

10 in

mm

and summer rainfall occur in the eastern grasslands


places where snowdrifts develop.

16 in

MONTH

Present-Day Environments and Climate Change

Fig. 3.7. Average maximum temperatures, 19812010. The


highest maximum temperatures correspond approximately
with the longest frost-free periods and the highest rates of
potential evapotranspiration. From Wyoming Water Resources

Temperature, Frost-Free Period, and Evaporation


The July mean daily high temperature ranges from 90F,
such as on the Great Plains and in the Bighorn Basin, to
less than 75F in the higher mountains.10 The July mean
annual low temperature is 55F on the plains and near
freezing in the alpine zone, where frost at night is common even in the summer (fig. 3.7). During a 24-hour
period, variation is most extreme in the alpine tundra,
where air temperature near the ground may be 80F or
more during the day and below freezing at night. Such
extreme variation is caused by rapid heating of the soil
during the day (stemming from reduced water vapor in
the atmosphere of the alpine zone), and, for the same
reason, rapid cooling at night. Although most plants
can acclimate to cold temperatures, many cannot adapt
to such extreme diurnal cycles. Notably, the whole
region is comparatively high and dry, with relatively

Data System, prepared with PRISM Climate Group data,


Oregon State University (http://www.prism.oregonstate.edu/
terms.phtml). Cartography by Christopher Nicholson.

rapid warming and cooling even in the lowlands, especially when there are few or no clouds.
Historically, the mean frost-free period varies from
125 days to fewer than 25 days, with the longest periods occurring in the Bighorn Basin, in a small area near
Riverton in the Wind River Basin, and on the eastern
plains (see fig. 3.7). Such areas have the largest amount
of cropland. Crop plants generally do not tolerate frost,
though many native plants do. Significantly, although
temperature has an important effect on the initiation of
plant growth in the spring, the termination of growth
is more often caused by drought, especially in the lowlands where precipitation is lower. Native plants can tolerate short growing seasons; most crop plants require
irrigation. Fallow dryland farming is an alternative,
which allows such crops as wheat to benefit from 2years
of soil moisture accumulation rather than 1 year.

33

34

Wyoming and the Rocky Mountain West

Topography influences growing season length and

chapter 8). The greatest potential for wind-generated

mean temperature by causing the flow of heavier cold

electricity is in the south, especially in the southeastern

air from mountain slopes to lower elevations, often

quarter of the state.14

along waterways. This airflow can create temperature


inversions, that is, warmer temperatures on mountain
slopes than in the adjacent lowlandsthe reverse of what

Soil Characteristics

is normally expected. Thus, the cities of Jackson and

Plant growth, crop production, and the carrying capaci-

Cheyenne are at about the same elevation6,209 feet

ties for wildlife and livestock depend on the nature of

and 6,053 feet, respectivelybut Jackson is much cooler

the soila matrix of organic matter, clay, silt, sand,

because of its location in Jackson Hole, which traps cold

and stones, teeming with microorganisms, inverte-

air flowing downward from the nearby Tetons.

brates, and roots, and mixed periodically by burrowing

11

One of the most significant effects of temperature is

mammals, such as prairie dogs, ground squirrels, and

on the potential rate of evapotranspirationevaporation

gophers. Indeed, the amount of energy consumed by

from soil and open water plus the amount of water

belowground organisms in grassland ecosystems can

evaporated on the interior of plant leaves and lost as

exceed the amount used by animals aboveground (see

vapor through the stomata. Higher temperature and

chapter 6). Every spoonful of soil contains millions of

lower humidity cause higher evapotranspiration, which

organisms, including bacteria, fungi, yeasts, protozo-

leads to summer drought stress for plants and animals.

ans, nematodes, and insect larvae. These mostly minus-

The potential evapotranspiration, which can be calcu-

cule organisms contribute immensely to the structure

lated from temperature and solar radiation, generally is

and fertility of the soil. Food webs belowground are

greater than annual precipitation in the plains, basins,

equally as complex and interesting as those above

and foothillsindicating dry conditions, because more

ground, though much more difficult to study.

water can be evaporated than usually is available. In con-

A multitude of different soils exists in the region,

trast, potential evapotranspiration in the mountains is

caused by the interactions over thousands of years of

less than annual precipitation, creating relatively moist

varying climate, topography, rock substrate (parent

conditions and the potential for streamflow. In the

material), and plant and animal life (see appendix B).

mountains, where more precipitation usually falls than

Sometimes the soils change gradually from one place

evaporates, there is the potential for loss of nutrients as

to another, forming a gradient from comparatively dry

they are dissolved in water moving through the soil pro-

grasslands to relatively moist meadows. In other places,

file, a process known as leaching. Only about 10 percent

the soils change abruptly because of abrupt changes

of Wyoming has a precipitation-evaporation ratio greater

in topography or the rock type from which soil min-

than one.12 Climate change surely will alter the precipi-

erals are derived. Some soils are shallow and not well

tation-evaporation ratio, which will lead to subsequent

developed, largely because of recent glaciation in the

changes in all ecosystemsincluding croplands.

mountains, or an abundance of erodible slopes and a

Topography and wind play an important role in


controlling water availability. South slopes are warmer

comparatively dry, cool climate not conducive to soil


formation.

than north slopes, enabling more evapotranspiration,

The physical and chemical characteristics of soils are

and wind accelerates evapotranspiration and leads

heavily influenced by the nature of the parent material,

to snow redistribution and sublimation. Areas blown

whether bedrock or transported sediments. Igneous and

free of snow are drier during the growing season; areas

metamorphic rocks (such as granite, basalt, quartzite,

where snow accumulates are wetter. Strong winds are

and rhyolite) are found primarily in the mountains.

common in Wyoming, especially in the south, where

They weather slowly, and consequently the soils are

the prevailing westerly winds are funneled through

usually shallow and coarse. Glacial moraines also lead

the Rocky Mountains at a low point in the Continental

to the development of coarse soils. Sandstones and

Divide (generally along Interstate 80).13 This wind pat-

limestones are resistant to weathering, often forming

tern has contributed to the development of a narrow

escarpments and ridges with shallow soils. Where suf-

band of sand dunes that is 100 miles long or more (see

ficient weathering has occurred, sandstones contribute

Present-Day Environments and Climate Change

to the formation of a coarse-textured soil with a high

aeration, and salinity. Infiltration is especially impor-

infiltration capacity for water. Sometimes the resulting

tant in semi-arid regions, where the potential for evapo-

sand is sorted by wind into extensive dune fields, such

ration is high due to relatively warm, dry air. Water that

as northeast of Rock Springs and between Shoshoni and

percolates rapidly to a depth of 2 inches or more evapo-

Casper (see figs. 1.5 and 9.1).

rates more slowly, remaining in the soil long enough

Fine-textured and deeper soils usually develop from

to be used by plants. Coarse-textured soils in semi-arid

more easily weathered shales, mudstones, and siltstones,

regions tend to have higher infiltration rates and con-

or from material transported by wind and water (such as

sequently higher rates of plant growtha principle

on much of the western Great Plains). Infiltration rates

referred to as the inverse texture effect. The name stems

are slower in fine-textured soils, but the water-holding

from the fact that, in eastern North Americawhere

capacity is higher. Such soils can be quite fertile,

the precipitation is generally higher than in the low-

though some are saline or high in toxic elements (such

lands of Rocky Mountain statescoarse-textured soils

as selenium). Saline playas develop where fine-textured

sustain less plant growth than those with fine texture

sediments accumulate in closed basins with no or little

(more silt and clay).15

outflow. Some shale-derived soils (for example, from the

The principle of the inverse texture effect can be

Mowry, Steele, Belle Fourche, Pierre, and Thermopolis

extended to rock outcrops, where the soil is thin or even

formations) are high in bentonitea clay that swells

nonexistent but where water is funneled into cracks.

and shrinks during wetting and drying cycles, which

Plant roots capable of growing through these fissures,

typically prevents the establishment of young plants.

such as the roots of mountain-mahogany and limber and

Similarly, soils high in soluble salts, gypsum (hydrated

ponderosa pine, probably have more water available to

calcium sulfate), and sodium may form crusts on the

them than would be the case for adjacent fine-textured,

surface that slow the rate of plant establishment (see

deeper soils. In fact, shrubs and trees throughout the

chapter 8).

region, which generally require more water than do

Predictably, soil characteristics change with eleva-

herbaceous plants, are often found on rock outcrops

tion. Organic matter content increases from arid sites in

or coarse soil, not on fine-textured soils, where water

the basins up to montane meadows, but then decreases

often evaporates before plants can use it (see fig. 3.3).

farther up into the alpine zone. Plant production is

Mountain-mahogany and other shrubs that grow in such

highest at mid-elevations, where more moisture is avail-

places are an important source of food for wildlife.

able but the climate is still not as cold as higher on the

Water-holding capacity is determined by soil tex-

mountain. Severe climates, such as in the alpine tun-

ture, soil depth, and the amount of organic matter.

dra, tend to restrict soil development as well as plant

Fine-textured soilswith a relatively high percentage of

growth.

silt, clay, and organic matterhold more water against

Salts and nutrients accumulate in the soil where

the flow of gravity than do coarse-textured soils. They

the precipitation-evaporation ratio is less than one, as

also enable more plant growth if the soil is saturated

would be expected during most years in intermountain

at least once during the year, such as often occurs near

basins and on the Great Plains. A layer of lime (calcium

snowdrifts. Abrupt transitions between grasslands and

carbonate) develops near the average depth of water

some shrublands are often associated with changes in

infiltration, forming a hardpan or caliche layer that

soil depth or soil texture (see figs. 17.9, 17.11, and 17.17),

may restrict root development. In contrast, a hardpan

with the more drought-tolerant grasses abundant on

does not develop in the mountains where the precipita-

shallower soils with a low water-holding capacity (see

tion-evaporation ratio is greater than one and the soils

chapter 7).

tend to be slightly to strongly acidic. With more water

Soil aeration is determined by texture and the

moving downward through the soil profile, the poten-

amount of water in the soil; it is important to plants

tial for nutrient leaching is higher.

because air spaces in the soil have the oxygen required

Five soil features seem especially critical in deter-

for root respiration. Oxygen is less readily available if

mining vegetation patterns in Wyoming and adjacent

the spaces are filled with water, which occurs more

states: infiltration rate, depth, water-holding capacity,

often in fine-textured soils or in wetlands. Greasewood,

35

Wyoming and the Rocky Mountain West

Blue grama
Gardner saltbush
Big sagebrush

20

80

30

clay

40

60

sandy
clay
clay loam

sandy clay
loam

20

60

silty clay
loam

70
80

loam
silt loam

10

loamy
sand
sand

100

50

silt

silty
clay

cla
y

10

90

50

40
30

Fig. 3.8. Surface soil texture has a strong influence on plant


distribution, as shown here for a grassblue gramaand
two species of shrubs, Gardner saltbush and big sagebrush.
Adapted from Nichols (1964).

100

70

36

90

80

90

sandy loam
70

60

silt
50

40

30

20

10

100

% sand

a common shrub around the edges of alkali flats, is well

The weathering of some rock types leads directly to

adapted to saturated soils with low oxygen, whereas big

the formation of soils with high salt concentrations,

sagebrush cannot tolerate standing water within 10 cm

especially Cretaceous marine shales, such as those in

of the surface.16 Well-drained, well-aerated soils are nec-

the Steele, Niobrara, Cody, and Pierre formations. Saline

essary for the survival of big sagebrush and the growth

soils often develop in basins where considerable water

of crops (fig. 3.8).

accumulates on the surface during wet periods and

Salinity, and other chemical traits such as alkalinity

where subsequent evaporation leads to the accumula-

and sodium concentration, depend on the composition

tion of salts on the surface, often forming white crusts.

of the soils parent material, the amount of water that

Soils derived from bentonite tend to have high con-

has evaporated from the soil over the centuries, and

centrations of sodium (because bentonite is composed

little or no potential for the leaching of salts from the

largely of montmorillonitic clay, which develops in

soil (that is, where the precipitation-evaporation ratio is

high-sodium sedimentary environments).

less than one).17 Plants that tolerate saline, alkaline, and


high-sodium soils are known as halophytes, for example,
greasewood, saltgrass, and the various species of salt-

Periodic Disturbances

bush. Less salt-tolerant species, such as big sagebrush,

The environment is usually described in terms of tem-

are sometimes found growing with halophytes but usu-

perature, wind, water availability, and soil characteris-

ally in areas where salt concentrations are lower (along

tics, but just as important are the various disturbances

drainages where snow accumulates and some of the

that have affected ecosystems for millennia. Examples

salts are dissolved and washed below the rooting zone;

include fires; windstorms; floods; burrowing; epidemics

see chapter 7).

of certain insects; and, in the alpine tundra, frost heav-

Present-Day Environments and Climate Change

ing of soils.18 Attempts to suppress disturbances may

re-established by sprouting. In the same way, grasslands

create an environment in which some native species

also recover in a short time, with little plant mortal-

cannot survive.

ity. In contrast, shrublands dominated by big sagebrush

Fire is a well-known natural disturbance that can

may not recover for 1020 years, simply because big

occur when there is sufficient fuel. Long before humans

sagebrush lacks the ability to produce new stems and

learned to use fire, lightning strikes provided the neces-

leaves (shoots) by sprouting from its root system.

sary ignition, even in grasslands. The number of light-

Windstorms constitute another periodic distur-

ning strikes to the ground is surprisingly large, often

bance. Individual trees are often toppled, creating small

several hundred during a single summer storm. Not all

canopy gaps important to the survival of some under-

of them cause fires, but once started, prior to the time of

story species. Occasionally, large numbers of trees may

active suppression, the fires probably burned for weeks

be blown down, as occurred in the Teton Wilderness in

or months over large areas. Native Americans frequently

August 1987 and the Routt National Forest in October

started fires, probably most often to facilitate their hunt-

1997. In a few minutes, the trees over large areas were

ing of big game. Whatever the source of ignition, early

broken or uprooted, creating a landscape reminiscent of

explorers often saw smoke on the horizon. Meriwether

that after the eruption of Mount St. Helens in Washing-

Lewis wrote about fire and the scarcity of timber along

ton; all other environmental factors are affected. Evi-

the Missouri River in 1805:

dence of windstorms is easy to find in mountain forests.

This want of timber is by no means attributable to


a deficiency in the soil to produce it, but owes its
origin to the ravages of the fires, which the natives
kindle in these plains at all seasons of the year. The
country on both sides of the river, except some of
its bottom lands . . . is one continued open plain,
in which no timber is to be seen except a few . . .
clumps of trees, which from their moist situation, or
the steep declivities of hills, are sheltered from the
effects of fire.19

Fire and windstorms are physical disturbances, but


biotic disturbances are just as important. Such burrowing animals as prairie dogs, pocket gophers, badgers,
and harvester ants create disturbances that often cause
the death of some organisms while providing a favorable environment for others (see chapter 6). The gaps
created in this way are individually small, but combined they modify the environment over large areas.
Furthermore, burrowing animals move about on the
landscape, subjecting a large proportion of the land
to burrowing over long periods. Such disturbances are

Fires could be annual events in some of the more pro-

sometimes a nuisance for livestock owners, but there

ductive grasslands, or they could be comparatively rare

are beneficial influences as wellbetter infiltration of

in some high-elevation forests. In the foothills, a fire

water in a semi-arid environment, better aeration, mix-

might be expected every 525 years. Modern-day sup-

ing of organic matter deeper into the soil, a higher level

pression of grassland fires creates a favorable environ-

of biological diversity, and generally more plant growth

ment for the establishment and growth of some trees

(see chapter 6). Though difficult to value monetarily, all

and shrubs.

are examples of ecosystem services.

All fires are conspicuous disturbances, but some are

Population explosions of certain insects also create

more severe than others. Crown fires in forests become

disturbances. The insects most commonly involved in

extremely hot and burn through the treetops, killing

the lowlands are grasshoppers and Mormon crickets,

most trees and abruptly changing the nature of the

which can reach densities of 525 per square foot in

ecosystem. In contrast, surface fires that burn slowly

grasslands and shrublands in some years, consuming

through the forest undergrowth often consume only

most of the herbaceous plant material in a short time.

the shoots of small plants and the organic material on

Similarly, several kinds of bark beetles have histori-

the forest floor; most mature trees survive, as do most

cally killed trees over large areas from time to time. In

shrubs, grasses, and forbs, because their roots and

the past 20 years the beetles have affected forests over

rhizomes remain alive. Small trees may be killed, but

a larger area than ever before observed, probably the

within a year or two, most of the herbaceous plants are

result of climate change. Aside from influencing spe-

37

38

Wyoming and the Rocky Mountain West

cies composition, insect outbreaks accelerate nutrient

of an area over a period of several decades or more. They

cycling and may affect forest flammability (see chapter

point out that unusually cold days will occur during the

11). Notably, these insects are native species and have

warmest month on record and that above-average pre-

coexisted with their hosts for millennia.

cipitation can fall in the middle of a 10-year drought.

Overall, an environmental analysis for any area must

Climatologists also emphasize that climate trends

consider the interaction of many factors that change

will not be the same everywhere. Even with an over-

either gradually or abruptly across landscapes and over

all trend of warming on Earth as a whole, some areas

time. No two areas have exactly the same environment;

will be cooler or wetter than normal. For this reason,

generalizations must be made with caution. This com-

the term climate change is preferred, rather than climate

plexity makes the environment difficult to conceptual-

warming or global warming. The rate of change seems

ize, but it is often helpful to focus on a few variables.

slow to some, with an increase of only a fraction of a

Key variables in the region include soil characteristics

degree every decade when warming does occur, but

(especially infiltration rates, water-holding capacity,

such small increases can lead to a higher probability

and salinity of the soil), water availability in the soil,

of summer drought if there is no concomitant increase

and seasonal and elevational temperature differences.

in precipitation. The result is slower tree growth, more

Periodic disturbances by fire, wind, burrowing animals,

frequent forest fires, and less streamflow from the

and a few insect species further modify an already

mountains. Also, as will be discussed, diseases and

diverse landscape mosaic. Regionally, different climates

insect epidemics can spread in ways that have not been

have caused the formation of grasslands, shrublands,

observed before. And, with warmer temperatures, the

forests, and tundra. Locally, the nature of the vegeta-

atmosphere can hold more moisture, thereby providing

tion is dependent on soil characteristics; topographic

the energy for extreme weather events. Flooding and

differences; snow drifting; and such disturbances as

snowstorms could be more severe.

fire, flooding, and windstorms.

Computer models generate specific predictions about


climate change and until recently have focused on large

Climate Change in Western States

areas. Some observers are skeptical of this approach,


but such models are required to synthesize all the geo

Climate change is nothing new. As described in the pre-

physical, atmospheric, and ecological data and processes

vious chapter, tropical forests once grew in Wyomings

that are involved. Modeling is used with confidence

Red Desert; millions of years later, glaciers extended

for weather forecasting and launching space vehicles.

down to the foothills. However, the prevailing notion

Warming projections based on models suggest that the

has been that such changes are so slow that they hardly

mean annual temperature in Wyoming and adjacent

mattered at human time scales. There were admonitions

states will increase by about 6F by 2100, perhaps 7F in

to be prepared for occasional droughts and cold winters,

the southwestern part of the statewarmer than at any

but few considered long-term trends. Now, during the

time in the past 10,000 years.21

past 3040 years, worldwide records reveal increases in

Computer modeling has been evolving for decades

air temperatures, and, not surprisingly, satellite images

as a tool for understanding complex systems, but such

have documented a steady, dramatic decline in Arctic sea

models should be tested when possible with actual mea-

ice. Similarly, repeat photos show how mountain glaciers

surements from specific places over long periods. Bryan

are retreating wherever they still occur. Rocky Mountain

Shuman, paleoecologist and professor of geology at the

glaciers are no exception.20

University of Wyoming, analyzed such data for Wyo-

Questions about the validity of warming trends are

ming.22 He used records from 30 weather stations for

often asked when record low temperatures occur, or

the period 18952007 to determine changes in three

about a drought after a year of heavy snowfall. Clima-

variables: annual average temperature, the high tem-

tologists emphasize that there is a difference between the

perature on the coldest day of the year, and winter pre-

weather that most people think about on a daily basis,

cipitation (NovemberMarch). His results, summarized

and climate, which is defined by the average conditions

in fig. 3.9, show an increase since 1978 in the frequency

Mean Annual
Temperature

2012
45.6F

Average
41.7F

1895
40.9F

Mean March
Temperature
2012
40.2F

Average
30.1F

1895
27.8F

1895
0.97 in / m

November March
Precipitation

Average
0.72 in/ m
2012
0.65 in / m

1895 1905 1915 1925 1935 1945 1955 1965 1975 1985 1995 2005

Year

Fig. 3.9. Summary of historic climate data for Wyoming. The


trends suggest that, during the past two or three decades, the
frequency of unusually warm years and warm March temperatures has increased and the amount of winter prec ipitation
per month has generally declined. The shaded area indicates
1 standard deviation of the averages for 18952012. The

data come from thirty long-time weather stations distributed throughout the state, all with records starting in 1895,
and were obtained from the National Oceanographic and
Atmospheric Administration (http://www.ncdc.noaa.gov/
cag/). Adapted from Shuman (2012); supplemented with
recent data.

40

Wyoming and the Rocky Mountain West

of unusually warm years and warm winters. This warm-

basins in Wyoming, there is no obvious trend in six of

ing has occurred throughout the state. Not surprisingly,

the basins, though a downward trend is suggested in

during this period some glaciers have disappeared and

four (see chapter 17). Changes in future rain and snow-

wildland fires and bark beetle epidemics have become

fall are likely to be variable across the state, but with

more widespread.23 Shuman confirmed that tempera-

warmer temperatures, evaporation will be greater. In a

ture trends were correlated with atmospheric green-

warmer climate, it is probable there will be less snow,

house gas concentrations. He could not find correlations

less irrigation water, less forage production, more acres

with solar irradiance, the Pacific Decadal Oscillation, El

burned in wildfires, and other related changes.

Nio/La Nia years, or changes in land use. Essentially

Although Earths climate has changed throughout

the same results have been found for Montana using

geologic time, the consensus of climate scientists is that

comparable data. 24

recent and ongoing climate change is occurring more

Could the temperature trends for Wyoming be

rapidly than before and that the best explanation for

part of a cycle when viewed over a longer period? To

these changes involves greenhouse gases produced by

address this question, Shuman examined changes in

human activities. 31 Also, even though most organisms

the 14,000-year-old pollen record preserved in lake

have adapted to previous episodes of climate change

sediments to infer temperature changes for Yellowstone

during the long history of life on Earth, through evolu-

National Park and the Bighorn Mountains (see chapter

tion or migrating to more favorable environments, biol-

2). These pollen records show the warming that caused

ogists wonder whether the current rate of change is too

glacial melting about 10,000 years ago, and also the cool

rapid for the evolution of many species to keep up. Local

and warm periods that have occurred since that time.

extinctions of species will accelerate if migration to

Shuman concluded that it is unlikely that warming dur-

areas having favorable climates is not possible. Almost

ing the past 40 or 50 years is part of a long cycle.25 He

everywhere, habitat has already been fragmented to the

also determined that recent temperatures in Yellow

point where conservation biologists are concerned that

stone National Park . . . have exceeded those recon-

some species will not have a pathway along which they

structed for the past >6000 years and are now similar

can move when migration is necessary (see chapter 18).

to the highest temperatures of the Holocene record.26

Also, potential destinations may already be fully occu-

Though less pronounced, his analysis of data for the

pied. Such issues are pertinent even if humans are not

Bighorn Mountains gave the same results.

the cause of the current episode of climate change.

In Wyoming, as elsewhere, changes in rain and

Slowing the rate of climate change will require sig-

snowfall are more variable and uncertain than tem-

nificant and controversial changes in the way industry

perature change. Shuman did find a decline in the

is powered, the way homes and workplaces are heated

frequency of wet winters. Moreover, based on climate

and cooled, and the modes of transportation that are

data since 1895, he found there were no winters for the

used. Progress has been made on using wind, solar

8-year period from 1999 to 2007 that were wetter than

power, and biofuels, and on the capture and sequestra-

the 83-year average from prior to 1978. Only 12 winters

tion of carbon dioxide, but some observers are skeptical

in the past 57 years have been wetter.27 Another study

about such alternatives. Proponents are convinced that

reported that the snowpack in the Rocky Mountains has

using energy more efficiently, combined with reducing

been dropping more in the past 30 years than in the

dependency on fossil fuels, is feasible and provides vari-

past few centuries, a decline reflected in data on stream-

ous benefits. These include opportunities for the eco-

flow and reservoir levels.28 According to the National

nomic growth associated with new industries that could

Weather Service, 2012 was the third-driest year in

help supply worldwide markets.

118 years.29 Still, the results are mixed. The National

Current rates of climate change are likely to adversely

Climate Assessment program projected that, by 2090,

affect some aspects of the economies of western states.

precipitation in Wyoming would be about the same or

For agriculture, the challenges will include finding

possibly 10 percent higher.30 When long-term precipi-

additional freshwater for more irrigation at a time when

tation records are examined for the 10 major drainage

water is even less available than it is now. This chal-

Present-Day Environments and Climate Change

lenge is pertinent to the Rocky Mountain region, where

long as the beetles have susceptible hosts of sufficient

so much freshwater is provided by mountain snowfall.

size and number. The extent of mountain forests will be

Most likely, there will be less ice and snow, on aver-

greatly reduced, with shifts in the dominant trees.34 A

age, than today, and summers will be longer and drier,

recent report on the effects of climate change in west-

because the snowpack will melt earlier in the yeara

ern states concluded that big sagebrush will become less

trend that has been under way for 30 years.32 More hay

common, and invasive plants, such as cheatgrass, knap-

for livestock will have to be imported if plant produc-

weed, and leafy spurge, will become more widespread.35

tion rates are down, making meat more expensive.

In this same report, U.S. Forest Service ecologist Megan

Municipalities and other industries will be implement-

Friggens and her associates concluded, By the end of

ing water conservation programs more frequently (see

the century, 55 percent of future landscapes in the West

chapter 18). As Steven Gray and Chamois Andersen

likely will have climates that are incompatible with the

concluded in their prognosis for the Rocky Mountain

vegetation types that now occur on those landscapes.36

region, The concern now is that climate change may


increase the impact of droughts, just as population

In sum, a 2007 analysis of climate trends by paleoecolo-

growth and other factors have greatly increased the vul-

gists John W. Williams and Stephen T. Jackson sug-

nerability of the West to any type of drought.33

gested that the characteristics of the emerging climate

In the case of forests and rangelands, fires are likely

in the Rocky Mountain region and elsewhere are novel,

to be larger and more severe than before because of more

having never occurred before. 37 How will western eco-

frequent extended droughts that increase flammability.

systems change? How can municipalities, agriculture,

Also, the spread of bark beetle epidemics throughout

and industries adapt? Such questions are considered in

western North America during the past decade has been

subsequent chapters. Understanding environments of

to a large degree attributed to warming (see chapters 11

the present is important, but anticipating the effects of

and 12). Most likely, such epidemics will continue as

climate change in the future now seems urgent.

41

This page intentionally left blank

Part Two Wetlands

This page intentionally left blank

Chapter 4

Riparian Landscapes

Streamside landscapes are highly valued real estate

ing banks, and new sediments are deposited as bars or

ecologically and economically. They account for a mere 2

on floodplains. Such developments lead to a continu-

or 3 percent of the land area, but an estimated 75 percent

ously varying riparian mosaic of woodlands, shrub-

of native animal species depend on them at some time

lands, meadows, marshes, sand bars, and mudflats.

during the year for food, water, and shelter. Animals also

Physical factors that determine the nature of stream-

move along the riparian corridors in both directions,

side vegetation include elevation, climate, the steep-

often passing through otherwise hostile environments

ness of the terrain, soil type, channel sinuosity, and

(fig. 4.1). In addition, riparian zones filter sediments and

the width-to-depth ratio of the stream. Other factors

nutrients, thereby improving water quality, and during

are beaver dams, grazing by large mammals, and water

the spring they store water in alluvial sediments that sus-

diversion projects for agriculture, municipalities, and

tain streamflow late in the summer. Domesticated ani-

industry. Like the vegetation, riparian soils are highly

mals often congregate in riparian zones, as do people.

variable across a floodplain and are strongly influenced

Cities, towns, roads, and cultivated fields usually are

by hydrologic factors. Commonly, there are overlapping

located there, and, not surprisingly, most riparian zones

lenses of gravel, sand, and silt, indicating the locations

at low elevations are privately owned.

and velocity of former stream channels.

From the smallest rivulet at high elevations to the

Separating riparian habitats from other kinds of

largest river in the lowlands, the riparian landscape can

wetlandsnamely, marshes, fens, wet meadows, and

be viewed as a continuum. Creeks that begin with the

playas (see next chapter)is commonly done by con-

melting of alpine snow cross the full range of climatic

sidering water movement. Rivers and streams have

conditions, changing from incised streams flowing

free-flowing, oxygen-rich surface water most of the

over cobbles and gravel to meandering rivers underlain

year. Exceptions are ephemeral streams that usually

by sand or mud. The environment changes gradually

flow only in the spring and early summer. In contrast,

along the waterway, except at occasional beaver dams,

nonriparian wetlands are usually associated with still

waterfalls, and rapids, or at entrances to canyons. Some

or slow-moving water, and the bottom sediments are

streams are long, originating in the mountains; others

more likely to be anaerobic for a longer time during the

are short, such as those that begin as seeps or springs

year. In general, all wetlands have soils that are wet long

in the lowlands. Creeks and rivers with headwaters at

enough during the year to enable the growth of plants

high elevations have peak flows in the spring, at the

that require large amounts of water. Such species are

time of snowmelt. New channels may form during

known as hydrophytes and are used to define the nature

floods. Also, old trees topple into the water along erod-

and extent of wetlands for legal purposes. 2

45

46

Wetlands

1
2
3
4
5
6

Missouri River Basin


7 Little Missouri
Yellowstone
8 Belle Fourche
Clarks Fork
Big Horn
9 Cheyenne
Little Big Horn 10 Niobrara
Tongue
11 North Platte
Powder
12 South Platte

Great Divide Basin


13 Great Divide

Great Basin
16 Bear

Colorado River Basin Snake River Basin


14 Little Snake
17 Snake
15 Green

Fig. 4.1. A network of creeks and rivers connects the landscapes of the region. Many smaller creeks are ephemeral,
lacking water during the driest part of the summer. The locations of Yellowstone Lake and the larger reservoirs are shown.
Drainages 1417 lie west of the continental divide. Drainage

13 is the Great Divide Basin (see fig. 1.2), which is formed by


a split in the continental divide and has no outlet. About 72
percent of Wyoming contributes to the Missouri River drainage. Drawing by Linda Marston.

From Rivulet to River

aspen, and balsam poplar). Associated shrubs include

As rivulets merge, a first-order stream develops with


the adjacent vegetation dominated by several species
of sedges, grasses, and willows (table 4.1). Meadows and
short willow shrublands (see fig. 2.5) become more conspicuous as two first-order streams come together, forming a second-order stream, and as two second-order
streams merge to form a third-order stream. Lower on
the mountain, such shrubs as thinleaf alder, water birch,
and tall willows become more conspicuous (figs. 4.2 and
4.3). In the foothills, trees are more common, primarily narrowleaf cottonwood, but also Engelmann spruce
and blue spruce3 (and in some places, lodgepole pine,

red osier dogwood, silver buffaloberry, thinleaf alder,


water birch, and several species of willowoften in
dense thickets (table 4.1). At still lower elevations, such
as in the Green River, Wind River, and Bighorn basins,
and on the Great Plains, plains cottonwood becomes
the most conspicuous tree (fig. 4.4).
Several studies have examined streamside vegetation
in the mountains.4 Collectively, the results suggest that
the patchy riparian mosaic is affected by various environmental factors, primarily duration of soil saturation,
soil depth and texture, frequency of flooding, depth to
the water table, oxygen availability for roots, duration
of snow cover, growing season length and temperature,

Table 4.1 Some characteristic plants of riparian plant communities in Wyominga

Common
name

Latin name

Broadleaf
woodland

Coniferbroadleaf
woodland

AlderTall
Short
conifer
willow
willow
woodland shrubland shrubland

Cinquefoil
silver
Greasesagebrush
wood
shrubland shrubland

Riparian
meadow

TREES
Engelmann
spruce

Picea
engelmannii

Blue spruce

Picea pungens

Lodgepole pine

Pinus contorta
var. latifolia

Boxelder

Acer negundo

Green ash

Fraxinus
pennsylvanica

Narrowleaf
cottonwood

Populus
angustifolia

Plains
cottonwood

Populus
deltoides

Peachleaf
willow

Salix
amygdaloides

Russian olive

Elaeagnus
angustifolia

Basin big
sagebrush

Artemisia
tridentata ssp.
tridentata

Silver
sagebrush
(mountain)

Artemisia cana
ssp. viscidula

Silver sagebrush (plains)

Artemisia cana
ssp. cana

Common
snowberry

Symphoricarpos
albus

Greasewood

Sarcobatus
vermiculatus

Rubber
rabbitbrush

Ericameria
nauseosa

Saltcedar

Tamarix
chinensis

Silverberry

Elaeagnus
commutata

Skunkbush
sumac

Rhus trilobata

SHRUBS

(continued)

Table 4.1 (continued)

Common
name

Latin name

AlderTall
Short
conifer
willow
willow
woodland shrubland shrubland

Cinquefoil
silver
Greasesagebrush
wood
shrubland shrubland

Broadleaf
woodland

Coniferbroadleaf
woodland

Riparian
meadow

Water birch

Betula
occidentalis

Woods rose

Rosa woodsii

Red osier
dogwood

Cornus sericea

Shrubby
cinquefoil

Dasiphora
fruticosa ssp.
floribunda

Silver
buffaloberry

Shepherdia
argentea

Thinleaf alder

Alnus incana

Booths willow

Salix boothii

Diamondleaf
willow

Salix planifolia

Drummonds
willow

Salix
drummondiana

Geyers willow

Salix geyeriana

Grayleaf
willow

Salix glauca

Narrowleaf
willow

Salix exigua

Park willow

Salix monticola

Shortfruit
willow

Salix
brachycarpa

Wolfs willow

Salix wolfii

Alkali
cordgrass

Spartina gracilis

Alkali sacaton

Sporobolus
airoides

Basin wildrye

Leymus cinereus

Bluejoint
reedgrass

Calamagrostis
canadensis

Foxtail barley

Hordeum
jubatum

Meadow barley

Hordeum
brachyantherum

GRASSES

Latin name

Broadleaf
woodland

Kentucky
bluegrass

Poa pratensis

Nuttalls
alkaligrass

Puccinellia
nuttalliana

Orchard grass

Dactylis
glomerata

Saltgrass

Distichlis
spicata

Slender
wheatgrass

Elymus
trachycaulus

Western
wheatgrass

Pascopyrum
smithii

Smooth brome

Bromus inermis

Timothy

Phleum pratense

Tufted
hairgrass

Deschampsia
caespitosa

Nebraska sedge

Carex
nebrascensis

Northwest
Territory
(beaked) sedge

Carex utriculata

Water sedge

Carex aquatilis

Horsetail

Equisetum spp.

Mountain
bluebell

Mertensia
ciliata

Rocky
Mountain iris

Iris
missouriensis

Common
name

AlderTall
Short
conifer
willow
willow
woodland shrubland shrubland

Cinquefoil
silver
Greasesagebrush
wood
shrubland shrubland

Coniferbroadleaf
woodland

Riparian
meadow

SEDGES

FORBS

A dash indicates that the species is absent or uncommon.

Fig. 4.2. (above) Mid-elevation riparian shrublands on Jack


Creek in the Sierra Madre. Common shrubs and trees include
Geyers willow, Booths willow, diamondleaf willow, alder,
shrubby cinquefoil, Engelmann spruce, subalpine fir, and
lodgepole pine. Herbaceous plants include elephanthead
lousewort, red and yellow Indian paintbrush, and various wetland sedges. The upland forests are dominated by lodgepole
pine and aspen at lower elevations and Engelmann spruce
and subalpine fir at higher elevations. Mountain big sagebrush, Idaho fescue, lupine, and desert Indian paintbrush are
common on the drier, south-facing slope in the foreground.
Elevation 8,500 feet. See also fig. 2.5.
Fig. 4.3. (left) Bebb and Geyers willow occur along creeks
in the foothills, forming a riparian greenbelt in otherwise
semi-arid landscapes, such as here along Little Sage Creek on
the north end of the Sierra Madre. Riparian zones often are
subject to heavy grazing and browsing by large herbivores,
causing some shrubs to have a pruned appearance. Shrublands dominated by Wyoming big sagebrush occur on the
upland. Elevation 7,500 feet.

Riparian Landscapes
Fig. 4.4. Rivers and creeks with
dependable streamflow often have
woodlands dominated by two species
of cottonwood. Narrowleaf cottonwood
grows in the higher basins, such as here
along the Green River; plains cottonwood is more common at lower elevations on the eastern plains and in the
Bighorn Basin. Wyoming big sagebrush
is taller and more dense in ravines
where snow accumulates, as illustrated
in this photo. Elevation 6,200 feet.

big game browsing, and ice damage in the spring. All are

states, the material in the terraces was commonly

affected by features of the terrain, such as valley width

deposited during floods at the time mountain glaciers

and orientation, size of the drainage basin, and stream

were melting, more than 10,000 years ago. At that time

gradient and sinuosity. Abrupt changes in riparian veg-

the volume of water was much greater, and it was laden

etation are typically associated with different riparian

with sediments that were deposited as glacial outwash.

terraces, and islands (fig. 4.5). Beaver dams, debris jams,


and roads can also have dramatic effects. Often, within

High

landforms, namely, point bars, channel bars, cut banks,


Incised
channel

short distances, the channel can change from braided to


The flatter terrain of some mountain meadows and
the lowlands causes streams to flow more slowly, leading to the development of alluvial soils over sometimes
broad floodplains. Channel meandering is typical at
lower elevations, except where rivers have cut canyons or where steeper gradients lead to the formation
of braided streams (see fig. 4.5). Freshly deposited sand
and gravel form bars, which provide excellent sites for
the establishment of some species of willow and, at

CHANNEL STABILITY

meandering or incised.
Cut
bank

Floodplain
Bar

Point bar

lower elevations, cottonwoods and willows. Across from


the point bars are eroding cut banks of the floodplain,

Channel
boundary

the water (fig. 4.6). Cut bank erosion on meandering rivers produces some of the sediments that form new bars
downstream. This natural erosion, along with flooding
and the occasional formation of oxbows, creates a riparian landscape that is constantly changing.
Above the floodplain are terraces, former floodplains that now are well above the river. In western

Low

where older trees and other plants eventually fall into

Braided

Meandering

CHANNEL TYPE

Straight

Fig. 4.5. Braided, meandering, and straight channels have


different levels of resistance to changes in channel location,
providing different environments for riparian plants. Bars are
formed by the deposition of gravel, sand, and silt.

51

52

Wetlands

solved from calcareous rocks and deposited on the former floodplain. Now, with periodic wetting, the salts are
slowly moved upward when the soils are wet and then
deposited on the surface as the water evaporates. Consequently, halophytes (salt-tolerant plants) are sometimes
found on terraces. Less salt-tolerant species occur on the
adjacent floodplain, where floodwaters periodically dissolve the salts and transport them downstream.
The most successful cottonwood regeneration occurs
along sand and gravel bars after major floods that de
posit floating seeds high on the bank, where the seedlings that develop are not likely to be washed away by
subsequent floods.6 If groundwater is close enough to
the surface to provide the water needed by the saplings,
a curvilinear band of trees develops that provides a living record of flooding patterns and channel migration
(figs. 4.74.9). As the channel moves across the flood-

Fig. 4.6.The Snake River in Grand Teton National Park, with


young trees and shrubs on point bars (on the right) and a cut
bank (on the left) with some trees about to fall into the river.
Douglas-fir is the dominant tree on the escarpment in the
foreground, and mountain big sagebrush steppe occurs on
the terracea former glacial outwash plainjust beyond the
river. Lodgepole pine forests occur on glacial moraines in the
background. See fig. 15.3 for a photograph taken approximately 90 to the left. Elevation 6,700 feet.

Now, with the glaciers gone and with different climatic


conditions, the rivers are smaller. They also have less
of a sediment load and, consequently, are more likely
to incise streambeds. Along some rivers there are two
or three pronounced terraces, each the result of a different period of glaciation and downcutting (see fig.
15.3).
Such substrate characteristics as texture, salinity, and
depth to water table are not uniform, causing patchiness in the vegetation. In Jackson Hole, for example,
mid-channel bars of sand and gravel are often occupied
by sandbar willow, fine-textured soils on the floodplain
have Booths willow, and the terraces regularly have
Geyers willow.5 Terraces at low elevations are more
likely to become saline because of salts that were dis-

Fig. 4.7. Band of plains cottonwood seedlings along the Bighorn River east of Lovell. Most of the seedlings die, but a few
survive if they are high enough on the bank to avoid being
washed away by subsequent floods and if groundwater remains
accessible. A curvilinear band of trees often develops, such as
those apparent in fig. 4.9. See Cooper et al. (1999) and Scott
et al. (1997) for estimates of flood size required for successful
cottonwood seedling establishment. Elevation 3,700 feet.

Riparian Landscapes
Fig. 4.8. Aerial view of riparian
woodlands along the Bighorn River
east of Lovell. Former channel locations and bands of trees are visible.
Most of the area shown is dominated
by plains cottonwood. The tree-ring
record indicates that, historically, a new
band of trees became established every
1020 years. The older bands tend to be
farther from the river (see figs. 4.9 and
4.10). Desert shrublands are found on
the adjacent uplands. The red in this
infrared image indicates an abundance
of chlorophyll in the riparian zone
where water is available longer into the
summer, in contrast to the adjacent
desert shrublands. Similar patterns are
found along other meandering rivers,
for example, the Powder River south of
Arvada. Image from the U.S. Department of Agriculture.

plain, the older trees die if their roots cannot keep up

a mile or more in width, and the immediate banks of

with the deepening water table.7 Those that survive are

the stream for 300 or 400 yards are covered with a thick

likely to fall into the river when the channel is again at

growth of cottonwood. Similarly, D. B. Sacket wrote in

their baseassuming beavers do not cut them before

1877: All along the Bighorn River [north of Thermopo-

that happens.

lis] . . . much fine, large cottonwood timber grows.9 In

The riparian zone at lower elevations is often wooded,


dominated by plains cottonwood. Other trees include

both places, plains cottonwood would have been the


dominant species.

peachleaf willow, ash, boxelder, lanceleaf cottonwood,

Notably, portions of some lowland rivers were histor-

and, rarely, American elm.8 Along the Wind River near

ically devoid of trees (fig. 4.10), such as along the Sweet-

Riverton, W. F. Raynolds wrote in 1860: The valley is

water and Laramie rivers. While camped along Crow

53

54

Wetlands
Fig. 4.9.This row of aging plains
cottonwood trees is not likely to be
replaced as the trees die. The riparian
meadow in this area, along Walker
Creek in Converse County, has Baltic
rush and various sedges. Cattails grow
in standing water, and introduced
plants (such as timothy and smooth
brome) are found nearby. Photo by Ken
Driese.

Creek, near Cheyenne, P. G. Cooke wrote in 1845, for

viding water late in the summer. However, if ground

forty miles we have seen but one treefive miles off

water drops to a depth where the large trees cannot

and not a bush or shrub.10 The near absence of trees in

obtain the water they need, woodlands give way to

some areas could have been caused by various factors,

meadows and shrublands. Extended droughts can lead

including frequent prairie fires, large mammal brows-

to the death of cottonwoods, just as shifting channels

ing, insufficient flooding for tree seedlings to become

and altered irrigation practices can.

established on point bars, or stable channels dominated

The riparian mosaic at the confluence of the Big-

by grasses and sedges that reduced the chance of tree

horn and Shoshone rivers in northern Wyoming illus-

seedling establishment (as discussed below). It seems

trates landscape patterns that can be found elsewhere

unlikely that the lack of seed dispersal is the explana-

(fig.4.11; see also fig. 4.8). Fertile soils and water avail-

tion, as many riparian species, such as cottonwoods,

ability for irrigation, combined with a relatively long

have seeds that are both buoyant and easily spread by

growing season, sustain one of the most important agri-

water and wind.

cultural regions in the state. Much of the native ripar-

Riparian meadows on the eastern plains, with no or

ian vegetation has been lost, but remnants persist in

only a few trees and shrubs, are commonly dominated

the Bighorn Canyon National Recreation Area. Plains

by Nebraska sedge, Baltic rush, prairie cordgrass, redtop

cottonwood woodland is the most conspicuous vegeta-

bentgrass, and plains silver sagebrush, plus introduced

tion type along the rivers. Associated species include

plants, such as Kentucky bluegrass, smooth brome,

peachleaf willow, silver buffaloberry, and the intro-

sweetclover, Canada thistle, hounds tongue, and tim

duced Russian olive. Various shrubs, forbs, and grasses

othy. In areas where salts accumulate in the soil, such as

are found as well (see table 4.1). Shrub and herbaceous

on terraces that are flooded less frequently, greasewood,

plant cover is low where tree density is high because

inland saltgrass, and alkali sacaton are common. Flood-

of inadequate light beneath the woodland canopy and

plain marshes with standing water commonly have

competition from trees for the available water.

dense growths of cattail (see chapter 5).


Trees have invaded some meadows and shrublands,

Interspersed in the woodlands at low elevations are


shrublands dominated by basin big sagebrush, plains

possibly because dams and reservoirs reduced the fre-

silver

quency of tree-damaging floods.11 Also, return flows

sandbar willow, skunkbush sumac, western snowberry,

from irrigated land may have favored the trees by pro-

Woods rose, chokecherry, silver buffaloberry, saltcedar,

sagebrush,

greasewood,

rubber

rabbitbrush,

Riparian Landscapes

Fig. 4.10.The Sweetwater River, looking northeast from Independence Rock in 2013. Note the lack of trees and shrubs, just
as in 1870, when a similar photo was taken by William Henry
Jackson (U.S. Geological Survery photo284; see K. L. Johnson

1987). Common plants include alkali sacaton, fringed sagewort,


American licorice, plains silver sagebrush, and rubber rabbitbrush. Note the elongated, well-vegetated point bar. Elevation of
the river is approximately 5,800 feet.

and a variety of forbs and grasses. Abrupt transitions

ian zones grow rapidly during spring and early sum-

from woodland to shrubland, together with abundant

mer, providing an abundance of highly flammable fuel

evidence of fire-scarred trees and logs, suggest that

by late summer. Shrublands dominated by skunkbush

fire has played an important role in converting older

sumac often replace the woodlands that are burned.

woodlands to shrublands. Cattail marshes occur in

Another important factor causing rapid change is

flooded oxbows, and meadows are found nearby, such as

flood control. Along the Bighorn River, Akashi found

in recently abandoned channels.

evidence of much less channel shifting after the


enlargement of Boysen Reservoir in 1952. Consequently,

Shifting Riparian Mosaics

fewer and smaller point bars were formed, resulting


in fewer sites for the establishment of cottonwood

The mosaic of some riparian zones is relatively stable,

seedlings.13 At the same time, trees continued to die from

because the creek or river is at the bottom of a V-shaped

old age, fire, beaver cutting, agricultural clearing, and

valley incised in bedrock. In contrast, riparian mosaics

increasing depth to the water table. Thus, the land area

on the alluvial soils of broad floodplains change rapidly

in old-growth cottonwood woodlands declined, a trend

(see figs. 4.8 and 4.11). Comparing a series of aerial

observed in other areas as well.14 Just as fire suppression

photographs taken over the Bighorn River, University

has led to changes in some mountain forests, flood

of Wyoming graduate student Yoshiko Akashi found

suppression has changed the riparian zone.15

that some of the woodlands, shrublands, and meadows

Notably, the shifting mosaic along the meander-

changed locations during a 50-year period.12 Fire was

ing Bighorn River is different from that on the nearby

a major factor causing the change. Plants in the ripar-

Shoshone River, a more rapid, braided river. Cotton-

55

Wetlands

Bar deposit
Sandbar meadow

Without
willow or
cottonwood

Cottonwood woodlands
Young (1-29 yrs)

Meadow

Basin big
sagebrush

Suspended sediment

Middle-aged (30-79 yrs)


Old (>80 yrs)

Fire
or
beaver
Fire
or
beaver

Saltcedar
willow

Annual
flooding

Saltcedar
Skunkbush sumac
Rabbitbrush
shrubland

Periodic
flooding
or fire

Bank erosion

Increasing distance from channel or height above river level

56

Periodic
flooding
or fire

Skunkbush sumac
Rabbitbrush
Snowberry
Rose shrubland

Prolonged inundation
?

?
Greasewood
shrubland

Fig. 4.11. Probable riparian succession on new point bars


along the meandering channel of the Bighorn River east of
Lovell. As indicated, succession has changed considerably
in this area since the introduction of saltcedar (also known
as tamarisk). With flood control, many point bars are not

suitable for the establishment of cottonwood seedlings. Therefore, the proportion of the landscape dominated by cotton
wood decreases as older trees die; shrublands have become
more common. Elevation 3,700 feet. Adapted from Akashi
(1988).

wood seedlings are observed more frequently along the

cottonwood in western Wyoming and eastern Idaho,

Shoshone, perhaps because favorable flooding is more

sandbars by sandbar willow, and mudflats by Booths

frequent there. By whatever means, habitat for seedling

willow.17 Cottonwood often invades stream margins,

establishment is created more often. Similarly, young

but as they age and the channel moves farther away,

narrowleaf cottonwoods are common along the free-

the trees die and give way to shrubland or grassland.

flowing Gros Ventre River in Jackson Hole.16

As noted, tree mortality in such places could be caused

The riparian landscapes along meandering and

by a dropping water table as the channel moves across

braided streams change so rapidly that it is difficult to

the floodplain, which enables the invasion of plants

think about succession in the traditional sense. Before

that are more tolerant of water stress and also possibly

the biota has a chance to form a relatively stable com-

benefit from more oxygen in the soil, such as red osier

munity, the river channel typically shifts to a new loca-

dogwood, Rocky Mountain juniper, silver buffaloberry,

tion or a fire occurs (because of the abundant fuel).

skunkbush sumac, and rubber rabbitbrush. The sup-

Still, several successional trends have been observed.

pression of disturbances along some rivers allows the

Commonly, cottonwoods and some willows are the pio-

development of self-perpetuating shrublands, and also

neer species, with gravel bars colonized by narrowleaf

woodlands with trees other than cottonwood.18 Along

Riparian Landscapes

the Snake River south of Jackson, blue spruce and red

Many people living along creeks and small rivers ben-

osier dogwood have formed a relatively stable commu-

efit from soil laid down by generations of beaver work-

nity. In eastern Wyoming, ash and boxelder form simi-

ing for thousands of years. Some have suggested that the

lar woodlands.

beaver is a keystone species, affecting ecosystem structure


and function far beyond what might be expected, con-

Bank Storage, Stream Hydrology, and Beaver

sidering animal size and population numbers. Others


have referred to beaver as bioengineers because of the

The plant growth on floodplains slows the movement of

structures and services they provide. To illustrate, sedi-

floodwater and thereby increases the potential for sedi-

mentation behind their dams provides clearer water

ment deposition when floods do occur. The accumu-

downstream, and greatly enhanced bank storage leads

lated sediments increase the amount of water stored in

to sustained streamflow throughout the year (figs. 4.12

the bank and floodplain, which, along with other fac-

and 4.13). A study in Colorado found that a valley with

tors, increases the probability of streamflow through-

an active beaver colony had eighteen times more water

out the year. However, maintaining bank storage has

storage in the spring and higher streamflow in late sum-

its costs in the form of water consumption by phreato-

mer than a comparable drainage where the beavers had

phytestrees and shrubs that live with their roots in

been removed.19 Higher late-summer streamflow bene-

groundwater much of the year. Every pound of plant

fits fish and wildlife as well as landowners. Beavers also

material produced may require as much as 400 pounds

thin woodlands by cutting trees, thereby stimulating

of water, because up to 99 percent of the water entering

the formation of new sprouts and the growth of under-

a plant through the root system is lost to the atmosphere

story plants. When the beavers are removed, the habitat

by transpiration. This water is required to keep stomata

can change quite rapidly. 20

open for the uptake of carbon dioxide and to maintain

Somewhat similar to beaver dams are the debris jams

turgid leaves for better light interception. But water

that commonly occur along wooded creeks (fig. 4.14).21

transpired by plants is not available for downstream

Debris jams are initiated most often by a tree that falls

uses. Such losses, along with seepage to groundwater

across a stream, or by floating wood and other materi-

aquifers and evaporation directly from water surfaces,

als that become lodged in the channel, creating a dam.

are known as conveyance losses. Water consumed by

Both kinds of dams improve water quality and benefit

riparian vegetation is sometimes viewed as unfortunate

some wildlife.22 However, a common reaction has been

in semi-arid environments, but it can also be viewed as

to remove obstructions to streamflow. The dams some-

a small price to pay for the benefits of increased forage

times appear messy, and they can inundate land that

and shade for livestock, erosion control, sedimentation

owners prefer not be flooded. Indeed, beavers become

during floods, biological diversity, better wildlife habi-

a nuisance when they build dams at road culverts or

tat, and sustained late-summer flowsall examples of

when they interfere with irrigation projects.

ecosystem services.

Beavers and their dams are much less common now

Few data are available to estimate quantitatively the

than in the early 1800s. Alluvial sediments originally

amount of water used by riparian vegetation, but two

deposited behind beaver dams have washed down-

generalities seem probable: first, streamflow depletion

stream in many areas. With fewer dams to dissipate the

by phreatophyte transpiration (that is, the proportion

energy of spring floodwaters, new gullies are created in

of streamflow used by plants) is higher in small streams

a year or two, sometimes leaving fence posts hanging

than in rivers; and second, depletion is higher during

from barbed wire in midair. As the channel deepens, the

late summer, when streamflow is low and transpiration

water table is lowered, and the soils on either side of the

is more rapid. Preliminary estimates suggest that reduc-

gully become dry earlier in the summer. The riparian

tions in streamflow by phreatophyte transpiration are

vegetation changes from willows and moist meadows,

less than 2 percent in June when streamflows are high,

which are relatively rare in the landscape and valuable

and up to about 50 percent in late August when the vol-

for livestock and wildlife, to the drier, already wide-

ume of flow is low.

spread vegetation of the adjacent upland. The riparian

57

Fig. 4.12.Centuries of beaver


activity have created a relatively flat riparian ecosystem
in what once might have
been a steep-sloping valley.
Common plants in this
area are water sedge; beaked
sedge; resin birch; and several
species of willows, including
diamondleaf, Wolfs, Geyers,
and Booths. Photo taken
along the upper reaches of
Lake Creek in the Medicine
Bow Mountains, southwest
of Dry Park. Lodgepole pine
dominates the forested upland, though many trees have
succumbed to the mountain
pine beetle. Elevation 9,100
feet. Persico and Meyer (2009)
discuss the variable effects of
beavers and climate on riparian landforms.

Fig. 4.13. Prior to the construction of this new beaver dam,


a deep gully existed along Muddy Creek near the town of
Baggs. The alluvial soils might have been deposited originally
behind previous beaver dams, which could have failed after
the beaver population was reduced by trapping in the 1800s.

It is possible that the current beaver population will continue


building dams higher and higher, eventually filling the gully.
Willow shrublands and meadows could become re-established
where there are now shrublands with greasewood and big
sagebrush. Elevation 6,000 feet.

Riparian Landscapes

Fig. 4.14.Trees falling across streams create debris jams


that provide habitat for fish and other organisms. Here,
Mullen Creek in the Medicine Bow Mountains is flanked by

Engelmann spruce, subalpine fir, and thinleaf alder. Elevation 8,500 feet.

zone narrows, and gully erosion continues as long as

be more successful if other construction materials were

the flows are rapid and unobstructed. Such flows pro-

provided. He cut several aspen trees in the distant foot-

vide less opportunity for the benefits of bank storage.

hills and transported them to near the transplanted bea-

Ironically, erosion along lowland streams is some-

ver. Several days later a dam was being constructed, and

times of sufficient concern that landowners resort to

within 2 years much of the gully behind the dam was

constructing dams themselves, which is far more costly

filled. With the rising water level, sagebrush and grease-

than the problems caused by beaver. As one Wyoming

wood became less common, while meadow grasses and

rancher noted, Beavers can be a pain in the neck when

willows became dominant. This experiment involved

you get too many in the wrong place. On the other

working with beavers and has been repeated success-

hand, how are we going to get that erosion stopped if

fully in several locations. Initially it was natural to look

we dont use beaver? They work pretty cheap, and Ive

for aspen to assist the beavers, but riparian ecologists

never seen a lazy one yet.23

have learned that even old tires wired together can

The role of beavers in preventing erosion and creat-

serve as a stabilizing framework around which beavers

ing habitat has long been recognized. It was therefore

can build new dams. Once the tires are in place, the

natural for land managers to consider transplanting

beavers finish and maintain the project using readily

beaver to severely eroded gullies. In one case, in south-

available shrubs.25

western Wyoming, the transplanted beavers had diffi-

Significantly, gullies are not a new landscape feature

culty building and maintaining dams made from the

created by human activities. As noted in chapter 2, gul-

plant material that was available, namely, greasewood

lies are frequently mentioned in journals written in the

and sagebrush. 24 A Bureau of Land Management biolo-

1800s. Some could have resulted from excessive bison

gist, Bruce Smith, wondered whether the beaver would

grazing or beaver trapping by the first fur traders, or

59

60

Wetlands

by erosion caused by wagon trains. Yet even the best

limbs and give them every night as the buffaloe [sic]

beaver construction sometimes fails, especially with

have entirely destroyed the grass throughout this part

exceptionally high spring floods. Beaver populations

of the country.28

could have been diminished for other reasons as well,

However, the explorers did not provide information

including disease and periods of drought. Whatever the

on the amount of time the bison spent in a particular

cause, the evidence suggests that, though gullies were

area. Without fences, bison would have wandered at

present prior to EuroAmericans, they are probably more

will, possibly grazing an area heavily but not returning

common now (see chapter 2).26

to the same place for a year or more. Also, the upland forage then, away from the rivers, might have been attrac-

Livestock, Reservoirs, and Irrigation


Riparian habitats have long been important to people
living in the semi-arid West. Beaver trapping and placer
mining for gold took place along creeks and rivers, and
abundant forage and irrigation water facilitated agricultural development. Most farms, towns, and cities
are located nearby, partly because of the availability
of water but also because of the aesthetic appeal of
riparian zones. Floods continued to occur, often causing great damage to riverside settlements, but usually
the accepted solution for the flood problem was dam
construction or channelization. Floodplain amenities kept people from moving to higher ground. Dams
not only prevented spring floods, they also provided
opportunities for waterpower, irrigated agriculture,
and recreation.
Today, more people than ever want access to the
resources provided by riparian landscapes. The result
is high land prices and increased concern about riparian management. Three topics are especially relevant
in this regard: livestock grazing, reservoir construction
and management, and the effects of different irrigation
systems.

Livestock

tive to bison more than to livestock. Today water tanks


and salt blocks are used to draw cattle to the upland, at
least for a portion of the year. But in some areas it seems
as though the only solution to conserving the riparian
habitat and streamwater quality is additional fencing or
labor-intensive herding.
One of the undesirable consequences of heavy grazing along creeks and rivers is bank erosion, which leads
to a decline in bank storage, water quality, and habitat.29 Such erosion can occur at any elevation, reversing a natural tendency, namely, sedimentation near the
banks where the flow is slower and a gradual narrowing
of the channel as sediments are stabilized by plants. The
deeper, narrower streams often have overhanging banks
that provide excellent fish cover (fig. 4.15). Hooved
animals, however, whether livestock or big game, can
destroy the banks, thereby reducing water storage capacity, widening stream channels, and increasing water
temperatures. 30 Consequently, fish habitat is degraded.
Several studies have documented these trends and the
fact that better livestock control improves fish production in small streams.31
Poor livestock management is not the only potential
cause of stream degradation. Bank erosion often occurs
with road construction and other developments, and
nutrients can enter the water by seepage from feedlots,
fertilized cropland, and inadequate sewage treatment

Domestic livestock concentrate where water, food,

systems. Such factors may be as serious for lakes and

and shade are in close proximity, sometimes leading

reservoirs as for the streams. A case in point is Flam-

to excessive grazing. Large numbers of bison could

ing Gorge Reservoir near the Wyoming-Utah border,

have had the same effect on streams as cattle or sheep.

where the upper part of the reservoir has algal blooms

Indeed, travelers in the early 1800s observed thou-

that can interfere with recreational activitiespossibly

sands of bison grazing along rivers.27 One of the earliest

because of nutrient additions to the Green River that

explorers, Osborne Russell, wrote about southern Mon-

cause undesirable changes to the aquatic ecosystem.

tana in 1835: The bottoms along these rivers are heav-

In the Greater Yellowstone Area, various investiga-

ily timbered with sweet cottonwood and our horses and

tors have concluded that willow and cottonwood were

mules are very fond of the bark which we strip from the

once more common when concentrations of elk on win

Riparian Landscapes

tion and other uses later in the summer. 34 However,


about 85 percent of the streamflow in Wyoming still
flows into neighboring states. 35 Further reservoir construction has been stalled by economic constraints,
environ
mental concerns, and binding agreements
with downstream states. The most direct negative
effects of reservoir construction are the loss through
inundation of a free-flowing stream and its associated
riparian habitat. Sacrifices are always necessary and
must be weighed against the obvious benefits.
One reason reservoirs are controversial is that their
utility can be short-lived because of sedimentation,
which continues regardless of location. Sedimentation frequently is accelerated by human activities that
increase erosion. Dealing with large, sediment-laden
reservoirs will be a problem for future generations. Beaver also cause sedimentation, but their activities are
restricted to small areas; the breaking of a beaver dam
is not a big issue.
The various purposes of reservoirs include flood control, irrigation, power generation, and recreation. Typically, water levels fluctuate considerablyhigh at the
beginning of the growing season to maximize the water
available for agriculture later in the summer, but low by
the end of the growing season because of withdrawals
Fig. 4.15.Excessive grazing and trampling by large animals
can convert cool, narrow creeks with overhanging banks
to warm, shallow creeks that are less suitable for fish. Also,
the sediments (black) that contribute to bank storage and
late summer flows are lost. Streams in poor condition can be
restored.

for irrigation. Reservoirs with hydroelectric plants, such


as those on the North Platte River, may have water
level drawdowns at times of high demand for electricity. When water levels are down, nearly barren shores
are created that sometimes become saline mudflats and
sources of dust clouds. Most native plants cannot tol-

ter ranges were much lower, due to predation by wolves


and American Indians (see chapter 15).32 Browse lines
on willows, juniper, and aspen are now commonly
observed throughout the region, sometimes the result
of browsing by native ungulates as well as livestock.33
Whether such browse lines are an indicator of excess
grazing is debatable (see fig. 10.16).

Reservoir Construction and Management

erate the water-level fluctuations that occur, although


such sites can become dominated by weedy species capable of rapid growth, such as dock, foxtail barley, goosefoot, knotweed, saltcedar, sowthistle, summercypress,
sumpweed, and sweetclover. In general, sedimentation,
water-level fluctuations, and the streamflow changes
associated with reservoirs inevitably disrupt the riparian ecosystem. The seriousness of these changes must
be judged after considering what is lost, what is gained,
and how long the benefits are likely to last.

Because of the states high elevation, Wyoming is part

Flood control levees also disrupt riparian land-

of an area known as the headwaters of the West. Snow-

scapes. The most prominent example in Wyoming is

fed streams flow into all major drainages (see fig. 4.1).

in Jackson Hole, where in 1957 the U.S. Army Corps of

There are 116 reservoirs with storage of 1,000 acre-

Engineers and other entities began channelizing the

feet or more that retain some of this water for irriga-

Snake River to protect valuable private property from

61

62

Wetlands

flooding at the base of the Teton Range. 36 The levees

crops or retained in the soil. But trade-offs are involved.

now extend for a distance of about 20 miles. Between

First, some of the spring flood water is diverted into

the levees, the river is narrower and flows deeper and

reservoirs, reducing the magnitude of beneficial early

more rapidly than before, causing more erosion down-

season flows. Second, much of the irrigation water

stream. Spawning habitat has been greatly diminished;

is transpired or evaporated from croplands, thereby

and without periodic floods and with less groundwater,

reducing the amount of water returned to groundwater

the floodplain flora and fauna have changed. Although

or streams. In most western states, evaporation from

enabling residential and commercial developments on

reservoirs and cropland (evapotranspiration) account

the floodplain, the levees have reduced the values of

for more than 90 percent of total water consumption. 38

riparian ecosystem services. Attempts have been made

Losses through evapotranspiration constitute a high

to restore some spawning grounds in tributaries to the

proportion of the total irrigation water in drainages

Snake, but at considerable cost.

where the amount of water used for irrigation is low,

Like levees, the dikes built across riparian zones for

or where the amount of water required for percolation

highways and railroads modify patterns of stream-

back to the stream or groundwater is greater than the

flow, thereby altering the riparian ecosystem. Ponding

amount of water applied. In other words, if water added

of water usually occurs upstream, which can be ben-

to a meadow or crop is less than the soil profile can

eficial for waterfowl, but it also elevates the potential

retain against the pull of gravity, then all or most of the

for flooding if culverts and bridges are not well engi-

irrigated water will be lost to evapotranspiration and

neered. Downstream from a roadbed, the floodplain

very little, if any, will flow back to the stream. Many

can become stabilized, depending on land use practices

factors are involved, including climate, the hydraulic

in the area and distance to the next streamflow obstruc-

conductivity and water-holding capacity of the soil, and

tion. Of course, the most dramatic effects of levees and

crop water requirements.

dikes occur during periods of high water. Notably, dikes

Irrigation by flooding has disadvantages as well

increase the width of riparian zones in some areas, as

as the potential advantage of sustained streamflows

does irrigation; levees reduce the width.

in some watersheds. For example, nutrient leaching


occurs more frequently with flooding, which leads to

Irrigation and Streamflow

fertilizer losses and concomitant nutrient enrichment


(eutrophication) of streams and groundwater. Moreover,

By whatever means, and whether the water comes from

salts that have accumulated in the soil over long

reservoirs, pumps, or streamflow diversions, irrigation

periods are dissolved and transported to the stream

occurs when resources are available to construct a con-

or groundwater, sometimes degrading surface and

veyance system and the soils and growing season are

groundwater with salinity that is too high for livestock,

adequate for the crops desired. In Wyoming, the area

cropland, or human uses.

with potential for irrigated croplands is small, only

Many farmers now have sprinkler or drip irrigation

about 23 percent of the state. However, the effects are

systems that are more efficient than traditional

more far-reaching than the amount of land involved.

irrigation, because the water is spread more uniformly,

One effect is that some streams with formerly yearlong

and a smaller portion of the available irrigation water

flows are now dry late in the summer because of with-

is transported to a particular field. The same amount

drawals upstream, creating ephemeral streams that are

of water can thus be applied to a larger amount of land,

much less valuable for fish, wildlife, and livestock. In

thereby reducing costs. There also is less potential for

1986 the Wyoming state legislature recognized fishing

the loss of fertilizer by leaching, and consequently

and recreation as beneficial uses of streamwater.37

stream eutrophication occurs more slowly. Increased

Significantly, flood irrigation can lead to more

crop production and lower fertilizer requirements

streamflow lower on the watershed later in the summer.

help pay for the required equipment and pumping

This happens because of return flows, that is, water

costs. However, although using irrigation water more

applied by flood irrigation that was not used by the

efficiently has considerable appeal, there is a greater

Riparian Landscapes

potential for salts to accumulate on the irrigated soil

ther. Some rivers were clogged with logs. After gold was

surface. Eventually, this salinization may reduce crop

discovered, feverish digging began adjacent to stream

growth. One solution is to flood the soil periodically,

channels. About that time, herds of thirsty livestock

but this alternative simply washes the salts into other

also trampled the shorelines, here and there, possibly in

bodies of water. The losses of one ecosystem are the

a way that was not typical of bison herds.

gains of another. Also, sprinkler irrigation produces

Thus, many riparian zones were destroyed or clearly

much less return flow. Less water is applied to a field,

disrupted by the end of the nineteenth century, at a

leaving less water that percolates into wetlands or

time when numerous exotic plants had been introduced

back to the stream. Consequently, with more efficient

to western statesplants that were well adapted to sur-

irrigation, water levels in all kinds of wetlands can be

vive in the relatively moist, fertile riparian habitat.41

lower in the fall.

Some were invasive, that is, they gradually displaced

39

native species.42 A few were introduced intentionally,

Invasive Plants, Climate Change, and the Future

because it was thought they might be economically


or aesthetically valuable; others arrived on their own,

The very first large-scale disturbance in the region by

such as Canada thistle, leafy spurge, yellow sweetclover,

EuroAmericans, on any landscape, occurred in ripar-

Kentucky bluegrass, and smooth brome. In retrospect,

ian zones, namely, commercial beaver trapping in the

many land managers now recognize that introduced,

early 1800s (see chapter 2). Prior to that time, disease

invasive plantsother than cropscan bring more

may have caused fluctuations in the number of beavers,

problems than benefits. Moreover, controlling such

but intensive trapping greatly reduced their population

plants is costly, if even possible.

sizes throughout North America. Many beaver dams

Two noteworthy invasive plants in the region are

surely failed, and in some areas new gullies formed. A

Russian olive and saltcedar (tamarisk). Russian olive

few decades later, some of the same waterways were sub-

was widely planted for windbreaks on the semi-arid

jected to a second round of human-caused disturbances

uplands and as an ornamental tree in towns. However,

associated with the logging industry and gold mining.40

it has invaded many riparian zones to the point that

Stream banks were cleared in many places, and sluice

many people think of it as naturalized (fig. 4.16)the

dams were constructed to create flash floods that car-

next thing to a native. Unlike cottonwood, it does not

ried thousands of logs, disrupting the floodplain fur-

require periodic floods for the establishment of new

Fig. 4.16. Russian olive, with gray-green


leaves, is an introduced invasive tree in
riparian zones at low elevations, such as
here along the Bighorn River south of
Thermopolis. Elevation 4,345 feet.

63

64

Wetlands

seedlings, enabling it to reproduce more rapidly than

and also whether such nutrient enrichment favors

cottonwood on the many streams in western states

the invasion of other exotic plants. Unusually high

where streamflow is regulated and floods are suppressed.

proportions of the herbaceous plants in many riparian

The fruits are commonly eaten and dispersed by birds.

zones consist of introduced species.

Most reproduction is from seedling establishment,

The other invasive plant, saltcedar, was introduced

though root sprouts do occur. Invasion is hastened in

to the region in the 1930s (fig. 4.17; see also fig. 4.11).45

some areas because beaver preferentially cut and eat

It grows mostly at low elevations, such as in the Bighorn

cottonwood sprouts, thereby creating a more favorable

Basin; the Wind River Basin; the lower parts of the

environment for Russian olive. And once established,

Green River, Powder River, and North Platte basins; and

Russian olive better tolerates the aridity that develops

northward into Montana. It has also been found above

with drawdown of the water table. As is typical for

7,000 feet in the Laramie Basin. With warming of the

invasive plants, part of the exotic trees success may

climate, saltcedar could become more widespread, as it

be attributable to the frequent preference of most

tends be more drought tolerant than some native shrubs

herbivores for native species.43

and trees.46 Both saltcedar and Russian olive grow

The alluvial soils of riparian zones generally have


sufficient nitrogen for plant growth. Nevertheless,

together on floodplains where the groundwater level is


dropping and upland plants are invading.

Russian olive has a symbiotic association with a

The seedlings of saltcedar tolerate saline soils better

group of fungi known as Actinobacteria, which form

than some native riparian species do.47 This tolerance

actinorhizae (analogous to mycorrhizae) that are capable

can be attributed partially to salt-secreting glands on

of nitrogen fixation. This adaptation may facilitate the

the leaves. Salts released by the glands are deposited on

exotic trees survival on the uplands, even though it

the soil from time to time by rainwater and also when

may not be critical on the floodplains. Various studies

the leaves are shed. Consequently, established plants

have shown that the soil under Russian olives has

can elevate surface soil salinity through litterfall.48

44

higher nitrogen levels than under cottonwood trees.

Saltcedar often grows so densely and uniformly over

Some have wondered whether nitrogen enrichment of

large areas that few native species survive, and it does

the streamwater is higher because of the Russian olive,

not support wildlife as well as the native vegetation. Like

Fig. 4.17. Saltcedar is an introduced


invasive shrub in riparian landscapes
that is becoming more common at low
elevations in Wyoming and southern
Montana. Another invasive plant, hoary
cress, is visible in the foreground. Photo
by Brian Mealor.

Riparian Landscapes

cottonwood and willows, saltcedar is a phreatophyte.

and dikes have been constructed to prevent flooding.52

However, it generally transpires more water than the

Along with such developments, the ecosystem services

native species do, most likely because it often produces

provided by riparian ecosystems have been diminished,

stands with more square feet of leaf area per square foot

creating costs that often are not consideredcosts

of ground area (the leaf area index).49

related to water quality, sustained streamflow, and wild-

Controlling the spread of both Russian olive and

life habitat.53

saltcedar, and many other invasive herbaceous plants

And now an added concern is warming of the

in the riparian zone, has been difficult. Herbicides

climate. Higher temperatures could lead to continued

typically kill some native species as well as the

decline in cottonwood abundance if floodplains become

invasives, and the chemicals are not recommended for

drier, floods are insufficient to enable new seedling

use in close proximity to waterways. Cutting the shrubs

establishment, or fires occur more frequently. One

or trees followed by the application of herbicide to

effect of a warmer climate already has been observed:

the root crown has been effective, though costly, and

earlier snowmelt, which increases the probability of

there has been some success with biological control.50

drought-caused mortality of riparian plants during a

A fundamental approach is to manage the river so

longer, warmer summer.54 Some rivers and streams that

that cottonwood, boxelder, and other native trees are

formerly flowed all year may become ephemeral, causing

favored. As the native trees grow taller, the saltcedar dies

dramatic changes to the riparian flora and faunanot

because of insufficient light.51 Of course, this approach

to mention municipalities, industry, and interstate

may require allowing floods and reducing the beaver

compacts.55 With a warmer climate, frost-sensitive

population, especially if cottonwood establishment and

invasive

growth is desired.

expand their ranges, and Russian olive could become

plantssuch

as

saltcedarprobably

will

more common, because it tolerates drought better


There is one clear conclusion for riparian landscapes:

than cottonwood can.56 Water shortages will promote

streamflow regulation, agriculture, irrigation, livestock

controversial initiatives, such as cloud seeding and

grazing, invasive plants, fire suppression, and other

additional trans-basin diversions, and also will reduce

human activities have created riparian habitats that

the availability of water for ecologically beneficial

are quite different from those of the 1800s. To varying

instream flows or floods. Sweeping generalizations for

degrees, alterations are occurring in both the lowlands

all riparian ecosystems are not possible, but its clear

and mountains, especially where roads and summer

that benefiting from riparian resources in the future

homes have been constructed in valley bottoms, large

will require a coordinated effort by policymakers,

herds of livestock or big game congregate, land is culti-

riparian scientists, fisheries biologists, engineers, and

vated, dams have created reservoirs and regulated flows,

land managers.

65

Marshes, Playa
Wetlands, Wet Meadows,
and Fens
Chapter 5

Nonriparian wetlands are those commonly referred to

lost at alarming rates, the U.S. Fish and Wildlife Service

as marshes, playa wetlands, wet meadows, and fens.

was charged in 1974 with creating a National Wetlands

Such places, along with the riparian habitats described

Inventory.2 Based on that work, it is known that wet-

in the previous chapter, are recognized for the benefits

lands now cover some 950,000 acres (384,000 hectares)

they provide by filtering waterborne pollutants, regu-

in Wyoming1.5 percent of the state. Considering

lating water flows, storing carbon, and enriching the

this small area and their ecological importance, wet-

diversity of plants and animals.1 For these reasons, and

lands are appropriately thought of as keystone habitats

because wetlands nationwide were being degraded or

(figs.5.15.3).

Fig. 5.1. Swamp Lake wetland


northwest of Cody is a mosaic
of cattail marsh, sedge marsh,
and fen. The fen is the largest
known peatland in Wyoming
and is classified as nutrient
rich. It receives water and
nutrients from the limestone
and dolomite of the Cathedral
Cliffs in the background.
Elevation 6,600 feet.

66

Marshes, Playa Wetlands, Wet Meadows, and Fens 67

Fig. 5.2. (above) Marsh in Hutton Lake National Wildlife


Refuge in the Laramie Basin. Hardstem bulrush is the most
common plant in this area and becomes abundant when
water levels are favorable. This marsh is dependent on water
from creeks and irrigation canals. Elevation 7,150 feet. Photo
by Rhonda Foley/U.S. Fish and Wildlife Service.
Fig. 5.3. (left) The yellow-headed blackbird is a summer resident of lowland marshes. Photo by Ken Driese.

Only 5 percent of Wyomings wetlands are permanently flooded, and two-thirds of them are dry by midto late summer nearly every year.3 At low elevations,
about half of the wetlands are sustained in part by irrigation waters, whether by flood irrigation, leakage from
irrigation canals, or runoff from irrigated fields.4 It is
likely that natural wetlands have a higher biological
diversity than those maintained by irrigation waters.
However, that has not been documented, and many
have water during years when natural wetlands are dry.
Of the 2,692 vascular plant species found in Wyoming, 12 percent are obligate wetland indicators, that
is, plants that cannot tolerate upland environments.
This proportion is surprisingly high, considering that
wetlands occupy such a small area in the state. An additional 13 percent of the plants are ranked as facultative
wetland indicators, that is, plants that usually grow

68

Wetlands

in wetlands but can tolerate moist upland habitats as

in invertebrates, providing a critical food source for

well. In all kinds of wetlands, plants in the sedge fam-

migrating waterfowl and shorebirds. Also, during low-

ily (Cyperaceae) are very common, with 47 percent of

water seasons, some plants and animals are favored that

the 174 species of that family in Wyoming classified as

are less tolerant of high water. Fluctuating water levels

obligate wetland indicators. The grass family (Poaceae)

enable both groups of species to survive. Water diver-

has more species (275) in the state than the sedge fam-

sions intended to keep wetlands filled to a constant

ily, but far fewer are confined to wetlands (8 percent).

depth may not be advisable, though the impulse to do

Among the shrubs, willows are especially common in

so is understandable if most nearby wetlands have been

mountain wetlands and in riparian zones, with 11 of the

drained, or if they have been severely altered in some

states 41 species (27 percent) found only in wetlands.

other way.

Other wetland shrubs include greasewood, resin birch,


and shrubby cinquefoil. Numerous sedges, grasses, and
other graminoids and forbs are found in all kinds of

Surviving in the Wetland Environment

wetlands (fig. 5.4). Trees, though, are largely riparian or

An abundance of water would seem to be an unalloyed

found on lake shores.

benefit in semi-arid climates, but prolonged soil satu-

Mosses and algae also can be an important compo-

ration creates oxygen deficiencies that many species

nent of wetland vegetation. A study of 17 fens in the

cannot tolerate.12 Those that do survive have various

Medicine Bow Mountains documented 30 species of

adaptations. To illustrate, microbial organisms derive

mosses, compared to 105 species of vascular plants. In

their energy in ways that do not require oxygen, using

Yellowstone National Park, a survey of 166 fens found

nitrates, sulfates, iron, manganese, and even carbon

44 species of mosses and 254 species of vascular plants.

dioxide instead.13 By changing these elements from oxi-

Notably, mosses in fens often cover as much of the

dized to reduced forms, they also create the chemical

ground surface as vascular plants. Algae are less con-

environment characteristic of hydric soils. For example,

spicuous but are an important source of food for her-

reduced iron and manganese are water soluble and are

bivorous aquatic invertebrates.6

leached from the soil, leaving white or pale gray hori-

Among the vertebrates, amphibians have the most

zons (known among soil scientists as gleyed horizons).

obvious affinity for wetlands, because the adults need

Similarly, much of the sulfur is reduced to hydrogen sul-

water for breeding and their tadpoles are confined to

fide, the marsh gas with an odor of rotten eggs. Reduced

water. Of the 12 species of salamanders, frogs, and

nitrogen gases are commonly lost from wetland soils,

toads that live in Wyoming, 10 are closely associated

so nitrogen can be a limiting factor for plant growth.

with water. Birds are the most conspicuous wetland

Similarly, carbonwhich exists in many organic forms

vertebrates, with 82 of Wyomings 434 species.8 Over

in hydric soilscan be lost as methane gas. Even so, the

half of those (44) are summer residents only.9 Of the

rate of carbon accumulation in undisturbed wetlands

118 species of mammals in Wyoming, approximately

usually exceeds the rate of loss.14

35 (30percent) use wetlands for more than a source of

Woody plants generally avoid the problems of anaer-

water.10 They include the beaver, muskrat, moose, river

obic soils by growing in microenvironments where oxy-

otter, raccoon, meadow jumping mouse, meadow and

gen is more readily available, such as raised hummocks

water voles, 3 species of skunk, 16 species of bats, and

and wetland edges, or on slopes where groundwater

8species of shrews.

moving through the soil supplies sufficient oxygen.

As noted, wetlands sometimes dry out, or at least

Mosses use the same strategy by growing on the wet-

the water level fluctuates dramatically. This variation

land surface. Some herbaceous vascular plants that are

seems unfortunate for many wetland species, but there

rooted deeper in the saturated soil obtain their oxy-

are benefits. Drying excludes predatory fish and allows

gen via aerenchyma, a spongy tissue in their stems and

the development of food webs composed entirely of

roots through which oxygen diffuses from the atmo-

invertebrates, including mollusks, crustaceans, and

sphere above the water to the respiring root cells. This

insects.11 Saline wetlands in the region can be rich

exchange can be more than passive diffusion, as some

Fig. 5.4. Seven grass-like plantsgraminoidsthat are common in some wetlands. The two species of cattail can be
distinguished by leaf width and the presence of a gap between

the spikes of male and female flowers. The male flowers of


both species persist for only a short time in early summer.
Drawings by Judy Knight.

70

Wetlands

wetland plants develop a pressure gradient that acceler-

Among both plants and invertebrates, the mix of spe-

ates the transport of oxygen to the roots.15 Plants with

cies in a wetland at a given time depends on the length

aerenchyma include buckbean, common reed, horse-

and timing of the preceding wet and dry periods.

tails, pond lily, sedges, and white marsh marigold.

Similarly, wetland vertebrates have evolved to opti-

Another problem for wetland plants is the toxicity

mize their chances of survival in seasonal wetlands.

of reduced sulfur, iron, and manganese compounds

Wetland birds, for example, fledge their young in the

produced by the soil microbes. Adapting to this stress

spring while marshes are usually flooded and emergent

involves the diffusion of oxygen from fine roots into the

cattails and bulrushes provide nest sites and cover. Simi-

millimeter-thick environment immediately adjacent to

larly, tadpoles develop and metamorphose into adults

roots, known as the rhizosphere. Oxidation in the rhi-

quickly, usually before the ponds dry out. Animals with

zosphere essentially immobilizes the toxic compounds

limited mobility, such as many amphibians, must be

and renders them virtually harmless. The oxidized iron

adapted for dry periods and winter. Spadefoot toads

of the rhizosphere remains long after the roots have

burrow into moist soil or occupy burrows dug by small

died and decayed, appearing in the soil as red or orange

mammals on the nearby upland, where they hibernate

streakinga diagnostic sign of hydric soils.

until the next favorable period.16

Vertebrates and terrestrial invertebrates avoid the


oxygen deficiency problem by living at the well-aerated
wetland surface. Aquatic invertebrates have the same

Marshes

array of adaptations as aquatic organisms in lakes and

For many, the word wetland brings to mind blackbirds,

streams, such as gills capable of extracting oxygen from

marsh wrens, muskrats, and a variety of emergent plants

the water.

that includes cattailsall characteristic of marshes.

As noted, two-thirds of Wyoming wetlands usually

Of the various wetland types, marshes are flooded to

dry out every year. In such wetlands, the microbes,

a greater depth and for longer periods. The substrate

plants, and animals must be adapted for drought as

of silt, clay, and organic matter is soft and, when wet,

well as soil saturation. Perennial plants that survive as

is commonly referred to as mud or muck. Hard-bottom

roots or rhizomes actually may benefit from short dry

marshes have more sand. The frequency, duration, and

periods, because some of the soil organic matter decom-

depth of flooding exert strong influences on the kinds

poses when exposed to the air, increasing the supply of

of plants and animals that are present. Many lakes do

nutrients during the next wet period. Another group of

not have marshes, primarily because their banks often

plants persists as seeds that germinate when water again

slope quickly into deep water, and wave and ice action

floods the wetland. A third group consists of plants

prevent plant establishment, or because water levels fluc-

that occupy the wetland during the dry period but are

tuate excessively, such as on the shorelines of reservoirs.17

absent during the wet period. These plants grow from

In the mountains, marshes tend to occur in small

seeds that persist in the soil during the wet period or

areas, such as closed depressions in glaciated terrain,

that blow into the dry wetland from nearby shoreline

also known as potholes (fig. 5.5). Other mountain

plants. Considering that some organisms benefit from

marshes are often associated with old beaver ponds and

dry or low water conditions, it is not surprising that

abandoned stream channels.18 Two common sedges

the diversity of plants and animals in wetlands can be

Northwest Territory sedge and water sedgeare the

greater if water levels fluctuate.

common dominants in shallow and moderately deep

Invertebrates also have strategies for living in sea-

standing waters in high-elevation marshes. Forbs occur

sonal wetlands. Immobile or slow-moving species, such

less frequently, but one can often find twinleaf bed-

as clams, survive dry periods buried in moist sediments,

straw, elephanthead lousewort, white marsh marigold,

either as adults or larvae, or they survive as eggs that

and pond lily, among others (table 5.1). Willows along

hatch during the next wet period. In contrast, aquatic

the margin of some mountain wetlands are commonly

invertebrates that metamorphose into flying adults

Booths willow, diamondleaf willow, Drummonds wil-

leave the drying wetland and fly to nearby water bodies.

low, Geyers willow, and Wolfs willow.

Marshes, Playa Wetlands, Wet Meadows, and Fens 71


Fig. 5.5. Pothole marsh on
the slope of Signal Mountain
in Grand Teton National
Park. Note the abundance
of pond lily with floating
leaves. Growing next to the
marsh is bluejoint reedgrass,
rough bentgrass (reddish), and
Northwest Territory sedge.
Elevation 7,000 feet.

At low elevations, common marsh plants include

gated fields. Small marshes sometimes occur in surpris-

Nebraska sedge, cattail, common spikerush, threesquare

ing places, such as in swales between sand dunes, where

bulrush, and hardstem bulrush. Forbs often are com-

water percolating through the porous dune encounters

mon and may include arumleaf arrowhead and water

relatively impervious sediments (see chapter 9).

knotweed. Marshes in the lowlands are found in aban-

In many marshes and playa wetlands, the vegetation

doned stream channels and backwater channels, but

forms concentric zones, each dominated by a differ-

they are especially common in irrigated areas, more

ent group of plants (figs. 5.6 and 5.7; see also fig. 5.5).

specifically, in the slowly moving waters of ditches and

Plants that need or can tolerate longer dry periods and

in depressions that receive water from canals and irri-

shallower water grow adjacent to the upland, whereas

Table 5.1. Some characteristic vascular plants found in or adjacent to marshes, wet meadows, fens, and playa wetlands in
Wyominga
Common name

Latin name

Marsh

Wet meadow

Fen

Playa wetland

GRASSES
Alkaligrass

Puccinellia spp.

Alkali cordgrass

Spartina gracilis

Alkali sacaton

Sporobolus airoides

Bluejoint reedgrass

Calamagrostis canadensis

LM

LM

LM

Northern reedgrass

Calamagrostis stricta

LM

Common reed

Phragmites australis

Saltgrass

Distichlis spicata

Short-awn meadow foxtail

Alopecurus aequalis

Tufted hairgrass

Deschampsia caespitosa

LM

LM

RUSHES, SEDGES, AND ARROWGRASS


Marsh arrowgrass

Triglochin palustris

Seaside arrowgrass

Triglochin maritima

Threesquare bulrush

Schoenoplectus pungens

Hardstem bulrush

Schoenoplectus acutus

Cottongrass

Eriophorum spp.

LM

Mountain (Baltic) rush

Juncus arcticus

LM

LM

Analogue sedge

Carex simulata

LM

Blister sedge

Carex vesicaria

LM

Buxbaums sedge

Carex buxbaumii

LM

Clustered field sedge

Carex praegracilis

LM

Mountain sedge

Carex scopulorum

Mud sedge

Carex limosa

Nebraska sedge

Carex nebrascensis

LM

LM

Northwest Territory (beaked) sedge

Carex utriculata

Sheep sedge

Carex illota

Water sedge

Carex aquatilis

Common spikerush

Eleocharis palustris

LM

LM

Few-flowered spikerush

Eleocharis quinqueflora

LM

SHRUBS
Alpine laurel

Kalmia microphylla

Alpine wintergreen

Gaultheria humifusa

Bog blueberry

Vaccinium uliginosum

Greasewood

Sarcobatus vermiculatus

Silver sagebrush (mountain)

Artemisia cana ssp. viscidula

Silver sagebrush (plains)

Artemisia cana ssp. cana

Purple marshlocks

Comarum palustre

LM

Common name

Latin name

Marsh

Wet meadow

Fen

Playa wetland

Resin birch

Betula glandulosa

Shrubby cinquefoil

Dasiphora fruticosa ssp.


floribunda

Booths willow

Salix boothii

Diamondleaf willow

Salix planifolia

Drummonds willow

Salix drummondiana

Geyers willow

Salix geyeriana

Sageleaf willow

Salix candida

Wolfs willow

Salix wolfii

LM

LM

Alkali buttercup

Ranunculus cymbalaria

LM

LM

Alpine leafybract aster

Symphyotrichum foliaceum

American globeflower

Trollius laxus

Arumleaf arrowhead

Sagittaria cuneata

Balsam groundsel

Packera paupercula

Broadleaf cattail

Typha latifolia

OTHER HERBACEOUS PLANTS

Narrowleaf cattail

Typha angustifolia

Buckbean

Menyanthes trifoliata

Elephanthead lousewort

Pedicularis groenlandica

English sundew

Drosera anglica

Field mint

Mentha arvensis

LM

LM

Horsetail

Equisetum spp.

LM

LM

Knotweed

Polygonum spp.

LM

Largeleaf avens

Geum macrophyllum

Marsh grass-of-Parnassus

Parnassia palustris

Narrowleaf bur-reed

Sparganium angustifolium

Neckweed

Veronica perigrina

LM

LM

Red swampfire

Salicornia rubra

Redpod stonecrop

Sedum rhodanthum

Redwool plantain

Plantago eriopoda

Sea milkwort

Glaux maritima

Seepweed

Sueda spp.

Spiral ditchgrass

Ruppia cirrhosa

Subalpine fleabane

Erigeron peregrinus

White marsh marigold

Caltha leptosepala

Riparian species are listed in table 4.1. A dash indicates that the species is absent or uncommon. L = present mostly at low elevations;

LM = present at both low and high elevations; M = present mostly in the mountains.

74

Wetlands

Fig. 5.6. Vegetation around desert playa wetlands changes


with distance from open water, probably because soil salinity
decreases and depth to water table increases. Salicornia, also
known as red swampfire, is a succulent that tolerates high
salinity.

and figs. 5.6, 5.7, and 8.4).20 Red swampfire, also known
as salicornia or saltwort, is a short, annual forb with
succulent leaves that grows on many salt flats and is
conspicuous when large numbers of the plant form a
scarlet band in late summer along the edge of white salt
deposits. Many salt flats are virtually unvegetated much

hydrophytes (such as cattails and hardstem bulrush)

of the time. In the Laramie Basin, for example, mudflats

dominate the deeper water. Different plants and animals are associated with each of the zones, an illustration of how patchy environments contribute to the
biological diversity of the landscape as a whole.
Most large marshes are found in areas that have
received water from irrigation or diversions. For ex
ample, cattails occur along the shore of Ocean Lake
in the Wind River Basin, which receives much of its
water supply from irrigation drains. In Goshen Hole
of eastern Wyoming, marshes with cattails and bulrushes occur on the edges of the lakes and ponds
enlarged or constructed on the Table Mountain and
Springer
/
Bump-Sullivan Wildlife Habitat Management units. In the Laramie Basin, the water levels of
Hutton Lake National Wildlife Refuge are managed by
U.S. Fish and Wildlife Service biologists to maintain
large stands of hardstem bulrush.

Playa Wetlands
Wetland desiccation is more likely to occur at low elevations, which sometimes leads to the formation of playa
wetlands (see figs. 5.6 and 5.7).19 When the surface water
evaporates completelymore frequently in playa wetlands than in marshesthe salts precipitate and become
sufficiently concentrated to exclude most plants. Only a
few halophytes tolerate such environments, including
greasewood, red swampfire, and saltgrass (see table 5.1

Fig. 5.7. Playa wetland in the fall, when most of the water is
evaporated and red swampfire is conspicuous on the white,
salt-crusted soil surface. The tan vegetation on the left is
dominated by saltgrass. Photo taken in the Laramie Basin at
an elevation of 7,000 feet.

Marshes, Playa Wetlands, Wet Meadows, and Fens 75


Fig. 5.8.Elongated wet
meadow south of Centennial,
with Sheep Mountain in the
background. This meadow
is defined by the presence
of shrubby cinquefoil, with
mountain big sagebrush and
scattered limber pine on the
upland. Elevation 8,100 feet.

with essentially no plants surround ponds with salinity

noids and forbs are the dominant plants; shrubby

greater than about 5 parts per 1,000. Ponds with less-

cinquefoil is common. (figs. 5.8 and 5.9).

saline waters support patches of marsh-like vegetation.

Sites that support wet meadows are diverse and occur

Playa wetlands are widespread on the plains and in

from the high mountains to the basins and plains.23

the basins of eastern and southern Wyoming. 21 Where

Commonly they are part of riparian landscapes. In the

water persists, the plants are mainly graminoids, such as

alpine zone, small meadows occur downslope from large

common spikerush, needle spikerush, threesquare bul-

snowbanks, where meltwater saturates the soil in spring

rush, mountain rush (also known as Baltic rush), and

and early summer. Broad glaciated valleys typically

meadow foxtail. Small forbs also are common, such as

support a mosaic of wet meadows and fens, with the

annual knotweeds. Around the depressions edge, where

meadows on the lower valley slopes and fens on wetter

facultative hydrophytes mix with upland plants, the

sites where groundwater discharges. Abandoned beaver

soil may lack the characteristic hydric features.

ponds are common sites for wet meadows, which replace


marshes as the ponds fill with sediments. In mountain

Wet Meadows

meadows, black alpine sedge and mountain sedge are


common at the highest elevations. Other species occur

Wet meadows develop where water tables are at or near

over a wide range of elevations, for example, tufted

the soil surface in the spring and early summer but

hairgrass, mountain rush, small-winged sedge, blister

where the water level drops below the surface as the

sedge, and water sedge. Common mountain forbs are

growing season progresses. Plant materials decompose

alpine leafybract aster, white marsh marigold, Ameri-

when water levels are low and oxygen is more readily

can globeflower, and largeleaf avens. Meadow shrubs

available, which prevents the accumulation of thick

include mountain silver sagebrush, shrubby cinquefoil,

layers of peat. Although wet meadows may not be rec-

and willowsespecially diamondleaf willow. Some

ognized as wetlands during part of the year, they have

meadows have substantial amounts of various mosses,

hydrophytes and the diagnostic hydric soils.22 Grami-

though they are more common in fens.

76

Wetlands

Fig. 5.9. Wet meadow in a bay on Upper Green River Lake,


in the foothills of the Wind River Mountains. The shrubs
are shrubby cinquefoil, and the single tree (foreground) is a
small lodgepole pine, both of which are growing where the
soils have more sand and are less anaerobic. The meadow is

dominated by graminoids, including blister sedge, bluejoint


reedgrass, fowl bluegrass, timber oatgrass, and tufted hairgrass. The forests are dominated by lodgepole pine, Engelmann spruce, and subalpine fir. Elevation 8,000 feet.

At low elevations, wet meadows commonly develop

and duration of high water. 25 In wetter areas the veg-

because of flood irrigation or leakage from irrigation

etation is composed of both obligate hydrophytes (like

ditches. Hayfields are an example, though they have

Nebraska sedge and water sedge) and facultative hydro-

mostly introduced species, such as Garrison creeping

phytes (such as bluejoint reedgrass and tufted hair-

foxtail, smooth brome, and orchardgrass. Native species

grass). The drier parts of meadows contain facultative

include tufted hairgrass, Nebraska sedge, clustered field

hydrophytes as well as upland plants.

sedge, common spikerush, and mountain rush. Wet

Because meadows remain moist during much of

meadows also occur on gently sloping valleys and broad

the growing season and are adjacent to rangelands on

flats without defined streams. Many low-elevation wet

the upland, plant growth is relatively high and they

meadows have areas with elevated salt concentrations

are often heavily grazed. 26 As in grasslands, prolonged

and halophytic species, such as alkali buttercup, alkali

grazing may cause changes in species composition,

cordgrass, alkali sacaton, Nevada bulrush, saltgrass, sea

from palatable to less palatable species. Also, excessive

milkwort, and seaside arrowgrass (see table 5.1). Wet

grazing can hasten drying and often causes the forma-

meadows can merge with salt flats, where few plants

tion of hummocks on the surface, a process known as

survive.24

pugging.27

Plants in wet meadows grow along a gradient in wet-

Wet meadows are desirable for real estate develop-

ness, from very wet sites that merge into marshes and

ments because of their proximity to surface water as well

fens, to drier sites that border upland vegetation. The

as to the upland, plus the availability of livestock forage.

important feature in this gradient may be the height

Federal laws restrict construction in wetlands because

Marshes, Playa Wetlands, Wet Meadows, and Fens 77


Fig. 5.10. Fens are common south of
Mammoth Hot Springs in Yellowstone
National Park. Northwest Territory
sedge and water sedge are abundant on
the usually wet organic soils; diamond
leaf willow, Wolfs willow, resin birch,
and shrubby cinquefoil are found along
the margin; wet meadows and lodgepole
pine occur on mineral soils that are
somewhat drier.

of their widely recognized ecological and hydrological

by precipitation that has not passed through the soil or

values. Thus, considerable controversy can arise about

rock and contains only the nutrients that settle on the

whether a tract of land is in fact a wet meadow or a

wetland in dust, rain, or snow.28 Both kinds of peatland

grassland or shrubland. Such disputes are settled based

store far more water and nutrients than an equal volume

on the abundance of obligate hydrophytes, the pres-

of mineral soil, though many of the nutrients are not

ence of hydric soils, and an understanding of the sites

available to plants because they are chemically bound

hydrology.

in organic material. Mosses do well in this environment,


probably because the reliable supply of surface water

Fens

minimizes water stress late in the summer and they are


able to obtain the small amount of nutrients required

Fens are rich in graminoids and forbs, but shrubs are

by such small plants. Peat in fens is composed mostly

more commonespecially willows and a shrub known

of slowly decomposing sedges and mosses. The true peat

as resin birch (figs. 5.10 and 5.11). The peat substrate is

mosssphagnumis found in some of them.29

saturated all year. Many people think of fens as bogs,

Fens are most common in glaciated mountain land-

but wetland ecologists have determined that all peat-

scapes, such as where potholes have formed and in val-

lands in the Rocky Mountains are fens. The distinction

leys. They also occur on slopes where a reliable source of

is based on nutrient availability and acidity. Bogs have

groundwater comes to the surface in springs and seeps,

low nutrient availability and are highly acidic (pH <

sometimes developing under the canopy of trees. 30

5). In contrast, fens are fed by groundwater containing

Most mountain fens are small, between 1 and 10 acres.

relatively high nutrient content and are either basic or

Fens in the alpine tundra can be associated with palsas,

mildly acidic (pH 58). Some of the plants are different

which are vegetated domes caused by frost heaving.31

as well.
The differences between bogs and fens are caused

Fens also occur in sagebrush steppe in the Green


River Basin and on the Sweetwater Plateau in west-

by their sources of water: fens are supported by waters

central Wyoming. The latter site, known as Ice Slough

that have moved through the soil or underlying rock,

(see fig. 5.11), is the subject of a historical marker west

from which nutrients are dissolved; bogs are supported

of Jeffrey City. It became famous because immigrants

78

Wetlands
Fig. 5.11.In the mid-1800s,
summer travelers along the
Oregon Trail would dig for
ice preserved in the peat of
this fen, known locally as Ice
Slough. Photo taken along
Highway 287, 40 miles east of
Lander. Elevation 6,500 feet.
Photo by Ken Driese.

along the Oregon Trail were happy to find ice preserved

stone National Park, adjacent to Wyoming Highway 296

in the insulating peat there. Core samples show that Ice

(see fig. 5.10).34 This large wetland complex lies on gra-

Slough is underlain by 510 feet of peat.

nitic bedrock and glacial deposits, but it benefits from

Fens sometimes merge with marshes as water depth

waterborne nutrients coming from the calcium- and

increases. In some places the fen community extends

magnesium-rich limestone and dolomite of the nearby

over the deeper water, with the peat forming a floating

Cathedral Cliffs to the south.35

mat. Eventually, the peat may cover the entire pond or

Fens share many plant species with shallow marshes

pothole, with the only evidence of a former pond being

and wet meadows (see table 5.1). In the wetter parts

the undulation of the surface when walked on, or the

of fens, Northwest Territory sedge and water sedge are

surprise of ones foot slipping into the water below.

ubiquitous across a broad elevation range, along with

Deciding where the fen ends and a shallow marsh or

other sedges. Some unusual graminoids grow in many

wet meadow begins, or even if a wetland contains a fen

fens, such as cottongrass and grass-of-Parnassus. Com-

at all, can be a challenge. Peat thickness can be used,

mon forbs are elephanthead lousewort, redpod stone-

but fens at high elevations commonly have thinner peat

crop, and white marsh marigold. Floating peat mats

than those at low elevations.32

usually are composed of analogue sedge, mud sedge,

Wetland scientists customarily identify different

and slender sedge. Along the edge can be found shrubs,

kinds of fens based on acidity and concentrations of

such as diamondleaf willow, mountain willow, Wolfs

calcium, magnesium, and other minerals in the water.

willow, small-leaved laurel, and resin birch.

At the acidic, infertile end of the gradient are poor fens,

Sometimes mosses are as conspicuous as vascular

similar to bogs. With decreasing acidity and increasing

plants. 36 Growing on the well-aerated surface of the

mineral concentrations, fens are labeled inter mediate,

peat, they respond readily to differences in surface

rich, and extremely rich. In general, the number of plant

water chemistry across small distances. In contrast,

species increases with nutrient availability. Most fens

vascular plants rooted in the peat are influenced more

in the Rocky Mountains lie on granite or other rela-

strongly by differences in climate, topography, and

tively nutrient-poor rocks and are intermediate.33 The

elevation, along with differences in water depth, water

most nutrient-rich fen identified in Wyoming is at

flow, and peat thickness.37 Different groups of plants are

Swamp Lake, in the Clarks Fork valley east of Yellow-

associated with floating mats, areas of standing water,

Marshes, Playa Wetlands, Wet Meadows, and Fens 79

hummocks, fen margins, and shallow channels with

as allelopathy (which has been difficult to demonstrate

running water.38

except in laboratory experiments).


Although both reed canarygrass and common reed

Management Issues and Future Challenges


Wetlands worldwide have long suffered degradation

are native to North America, they are also native to


Europe. Such a widespread distribution can be attributed to great genetic variation within the species. At

and outright destruction caused by myriad factors,

some point during the past century, seeds from Euro-

including the perceived need for drainage and water

pean populations of both species were introduced to

diversion, the accumulation of excess nutrients and

North America, where the resulting plants interbred

contaminants, and excessive grazing by livestock (see

with native populations. As it turns out, the hybrids

chapter 4). In Wyoming, 3540 percent of the original

appear to be more aggressive invaders than the native

39

wetlands have been lost, and elevated levels of ground-

populations. Thus, where reed canarygrass and com-

water contaminants are sometimes present.40 Generally,

mon reed appear to be invading wetlands, most likely

mountain wetlands have been less affected because

the plants involved are EuroAmerican hybrids. Slight

water is more readily available and the duration of graz-

genetic changes can greatly improve a plants ability to

ing is shorter.41 Also, pollutants are less widespread in

compete with its neighbors.

the mountains. Peat mining has severely disrupted

Invasive hybrids have also formed in cattails. Two

some fens in the mountains of Colorado,42 but that has

native species occur throughout much of North Amer-

not taken place in Wyoming.

icabroadleaf cattail and narrowleaf cattail. Until

Invasive plants have become a severe problem in

recently, the narrow-leaved species was so rare in mid-

many wetlands.43 Ecologists think of wetlands as land-

western and western states that some botanists, after

scape sinks, that is, low places on the landscape where

observing how its range was expanding, came to think

water, nutrients, sediments, and salts accumulate, and

of it as an introduced invasive plant.46 Indeed, narrow

where water levels fluctuateall conditions that favor

leaf cattail has expanded its range. More important,

weedy plants. Various Eurasian species are common

though, the two native speciesnow growing together

in Wyomings wet meadows, namely, Canada thistle,

more commonly than ever beforeare hybridizing. The

creeping bent grass, creeping foxtail, Kentucky blue-

sterile offspring, known as hybrid cattail, spread rapidly

grass, and timothy.

by rhizomes and are highly invasive.47 Very likely, Euro-

Some invasive wetland plants are now thought of

pean genes are a contributing factor.

as natives. Kentucky bluegrass is an example, as it has

An especially worrisome invasive plant in marshes

been in North America long enough to be viewed as

is purple loosestrife, a colorful forb that grows 3 or 4

naturalized. In contrast, some species that are veritable

feet tall. Though still rare in Wyoming, this European

natives behave like invasive species in altered wetlands.

species has overwhelmed native wetland plants in

For example, reed canarygrass expands rapidly in wet-

midwestern and eastern states. Seasonal wetlands are

lands created by irrigation. Such habitats could have

most vulnerable to invasion, because loosestrife seed-

been invaded by introduced species, but native plants

lings establish on bare sediments during dry periods.

often colonize the irrigated lands first. The species does

Preventing the spread of localized invasives like purple

provide good forage, but it tends to reduce the plant

loosestrife is generally more feasible than trying to

diversity of the areas where it becomes the dominant

reduce large populations once they are well established.

plant.44 Another invasive native, common reed (also

Also, early control is less damaging to the environment.

known as phragmites), grows to more than 6 feet tall

Biological control of purple loosestrife shows some

and is rapidly invading wetlands, especially at lower

promise.

This invader is thought

In addition to invasive species and the loss or degra-

to benefit from chemicals produced and released by

dation of wetlands, there are now concerns about the

the plants that suppress the growth of other species,

effects of climate change. Warmer, drier conditions

thereby facilitating its spreada phenomenon known

likely will result in shallower water, causing seasonal

elevations in adjacent states.

45

80

Wetlands

wetlands to dry up for longer periods and, for example,

ing through the soil from ditches and irrigated fields.53

converting marshes to playa wetlands.48 Sediments in

Surface inflows tend to be less saline, enabling different

Great Plains wetlands indicate that such changes have

wetland communities than where subsurface flows are

occurred in response to climate changes throughout pre-

the primary source.

history. During long dry periods, obligate hydrophytes


became less abundant and facultative hydrophytes

In sum, recognition of wetland values, their vulnerabil-

more abundant. In fens, increased decomposition rates

ity to degradation, and their poorly protected status has

resulting from warming and less flooding could cause

led state and federal agenciesalong with motivated

peat deposits to thin.49 Many kinds of wildlife and other

private organizations and landownersto promote

organisms will be affected adversely as the area in wet-

wetland conservation. One federal initiative is the Wet-

lands is reduced (see chapter 4). Ecologist Holly Cope-

land Reserve Program, administered by the U.S. Depart-

50

land and her associates conducted a recent assessment

ment of Agricultures Natural Resources Conservation

of Wyoming wetlands, concluding that climate change

Service. Landowners are paid to protect wetlands. Simi-

is most likely to affect wetlands at low elevationsthe

larly, the Wyoming Wetlands Conservation Strategy

same wetlands that are vulnerable to rural subdivisions

identifies wetlands where collaborative conservation

and oil and gas developments.51

management is needed and will work. 54 Unfortunately,

Agriculture is usually considered a cause of degrada-

wetlands often occur in places considered desirable for

tion in wetlands. However, as noted previously, many

new construction. When this happens, the prevailing

wetlands at lower elevations have been created or aug-

attitude is to construct new wetlands, to compensate

mented by irrigation projects. In Star Valley; Goshen

for those that are lost. Although commendable, this no-

Hole; and the Bighorn, Laramie, and Wind River basins,

net-loss policy sanctions the destruction of wetlands in

more than 50 percent of the wetlands receive irrigation

one place if a new wetland of comparable area is created

waters.

In the Laramie Basin of southeastern Wyo-

elsewhere. Many wetland scientists doubt that newly

ming, 30 percent of the inflows to wetlands were from

created wetlands provide the same benefits as those that

irrigation ditches, and 35 percent were waters percolat-

have been lost.55

52

Part Three Plains and

Intermountain Basins

This page intentionally left blank

Chapter 6

Grasslands

Grasslands are characteristic of the plains east of

scarlet globemallow, and various species of milkvetch

the Rocky Mountains and in several intermountain

and locoweed. Foothill grasslands at slightly higher

basins (see fig. 1.5). Some have been plowed for crop

elevations may have some of the same species, but

productionmostly winter wheatbut large areas of

others are more typical, such as bluebunch wheatgrass.

native vegetation can still be found (fig. 6.1). Prong-

Some plants are good indicators of soil characteristics.

horn are common. Diversity is added to the landscape

For example, sandy soils often have an abundance of

by scattered woodlands of ponderosa pine, limber pine,

Indian ricegrass, prairie sandreed, sand dropseed, sand

and juniper on outcrops of resistant bedrock, and by

sagebrush, and yucca, whereas saline soils have alkali

sand dunes and badlands (see chapters 8 and 9).

sacaton, fourwing saltbush, greasewood, and inland

Most grasslands in Wyoming have been classified

saltgrass. The most common soil types of upland

into two categories. One, known as shortgrass prairie,

grasslands, known as Aridisols, have accumulations of

is found only in the southeast corner of the state and

sand, silt, clay, and organic matter along with varying

has buffalo grass and blue grama as co-dominants. The

amounts of lime (calcium carbonate), gypsum (hydrated

other, mixed-grass prairie, covers a larger area (about

calcium sulfate), or various soluble salts. In swales and

17 percent), extending northward into Montana. Usu-

along drainages, where plant growth is comparatively

ally mixed-grass prairie does not have buffalo grass.

high for a longer period of time each year, the soils have

Sometimes the grasses in both types of grassland grow

more organic matter and are known as Mollisols (see

in small clumps and are known as bunch grasses, most

appendix B).

notably blue grama, junegrass, Sandberg bluegrass, and

Some grasslands mergegradually or abruptlyinto

Indian ricegrass. Others, such as western wheatgrass and

shrublands dominated by sagebrush or woodlands dom-

prairie sandreed, have rhizomes and are sod formers. 2

inated by juniper, limber pine, or ponderosa pine. Silver

Grasses provide most of the biomass in grasslands, but

sagebrush is common on sandy soils, in swales, and

forbs, succulents, and small shrubs provide most of the

along drainages; Wyoming big sagebrush is common

plant diversity. 3

in the intermountain basins (see chapter 7). Sand sage-

Mixed-grass prairie often has 50 or more differ-

brush is common on dunes, such as near Torrington,

ent kinds of plants in an area as small as 2 or 3 acres.

though it is less widespread in Wyoming than in east-

Common species include blue grama, Indian ricegrass,

ern Colorado. Around the world, shrubs are invading

junegrass, needle-and-thread grass, Sandberg bluegrass,

grasslands in a surprising number of places, a trend that

western wheatgrass, needleleaf sedge, threadleaf sedge,

has been attributed to various combinations of factors,

fringed sagewort, Hoods phlox, pricklypear cactus,

including less frequent fires, excessive grazing by large

83

84

Plains and Intermountain Basins

Fig. 6.1. Ponderosa pine grows on ridgetops adjacent to this


mixed-grass prairie on the western Great Plains; cottonwood
sometimes grows along creeks. The tan grass in the foreground is little bluestem. Other common plants are blue

grama, Sandberg bluegrass, western wheatgrass, junegrass,


needle-and-thread grass, and fringed sagewort. This grassland is near the Black Hills, at an elevation of about
4,500feet.

herbivores, nitrogen enrichment from atmospheric fall-

times causes the formation of patches that are one to

out in the form of dust and rain, and elevated levels of

several yards across (such as those for blue grama, prairie

carbon dioxide in the atmosphere.

sandreed, and yucca). Another cause of patchiness is the

Wherever grasslands occur, plant species composi-

burrowing of animals, such as badgers, ground squir-

tion varies with changes in topographic position, such

rels, harvester ants, pocket gophers, and prairie dogs.

as from hilltops to valley bottoms. Valley bottom soils

Their small disturbances favor some plants over others.

are deeper, finer-textured, moister, and more fertile than

Also, fairy rings may developan interesting but

soils on hilltops. Topography also affects snow drifting,

poorly understood phenomenon thought to be caused

with more snow accumulating in ravines and on leeward

by saprophytic fungi (fig. 6.2).5

slopes, causing them to be wetter than the surrounding


area. Depressions or playas with little or no drainage are
common, and typically they have saline or alkaline soils.

Surviving in the Grassland Environment

Differences in soil depth, salinity, and texture cause con-

Most of the grassland ecosystem is hidden in the

siderable variation in what sometimes appears as a uni-

soil, where 75 percent or more of the plant biomass

form vegetation covermonotonous during mid-day but

is found and, not surprisingly, most of the herbi-

inspiring when the sun is low in the sky.

vores also (fig. 6.3).6 Extensive root systems, required

Other patterns occur on a finer scale. For example,

to obtain water for the relatively small aboveground

cloning from root crowns, rhizomes, and stolons some-

shoots in this semi-arid environment, provide food

Grasslands

nial and herbaceous, having their perennating buds at


or just below the soil surface. This growth form is the
result of several factors: first, the semi-arid nature of
mixed-grass prairie limits the production of woody
stems aboveground. Second, and perhaps most important, is the frequency of fires in grasslands. Plants with
buds at or near the soil surface are less likely to be
killed, because the soil surface temperature during a
fire is relatively low. Most buds survive, and only the
herbaceous stems and leaves are burned. After the fire,
the buds sprout and grow rapidly. Third, plants with
their buds near the soil are well adapted to tolerate
grazing by large herbivores, which 10,000 years ago
included the camel and horse as well as bison, elk, and
pronghorn. Buds close to the ground, or buried in the
soil, are less likely to be eaten by the herbivores. Only
the easily regenerated stems and leaves are consumed
(fig. 6.4).
The ability to replace leaves and stems that have
been eaten or burned depends on the presence of special meristemstissues capable of cell division and
new growthand the capacity for energy storage in
the undamaged part of the plant belowground. The
meristems are contained in dormant buds on rhizomes
and root crowns, which are stimulated by herbivory to
Fig. 6.2. An aerial view of grasslands in the Laramie Basin
showing fairy rings that are about 30 feet in diameter, created by fungi that grow outward. A flush of nutrients is made
available as the fungi die and decompose, just inward from
the zone of major fungal activity, thereby stimulating plant
growth. Elevation 7,300 feet.

produce new stems and leaves. Another way that grass


leaves can be replaced is through intercalary meristems,
which are tissues capable of cell division that occur near
the bases of the leaves and some stems. As grass leaves
are eaten, new leaf tissue is produced by the leaf itself.
No buds are involved.7
Regrowth, whether from buds or intercalary meri-

for the largest group of herbivores in the ecosystem

stems, requires energy stored as carbohydrates in roots,

invertebrates. As the roots and other forms of life die,

rhizomes, bulbs, and corms. Carbohydrates are pro-

they become part of the soils organic matter. Some

duced by photosynthesis aboveground but are mostly

plants have deep roots for extracting moisture from

moved belowground soon thereafter. Considering that

throughout the soil profile, whereas others have shal-

75 percent or more of grassland plant biomass is below-

low roots to take advantage of light showers that wet

ground, there is a large amount of energy available to

only the surface soil. Generally, about 70 percent of

keep the plant alive and enable the production of new

the roots are within the top 4 inches.

leaves and stems.8 Of course, plants die from a lack of

Throughout the western Great Plains, grassland

energy if fire, herbivory, or drought are too frequent to

plants have experienced extended periods of drought,

allow for the replacement of the stored energy spent in

recurring fire, and grazing by herbivores. These envi-

previous episodes of regrowth.

ronmental factors led to a community composed

Most grassland plants have characteristics of xero-

largely of grasses but with a substantial number of

phytes, that is, plants adapted to dry environments.

sedges and forbs. Notably, most of the plants are peren-

Typically, they have small or narrow leaves, and some

85

86

Plains and Intermountain Basins

Fig. 6.3. Most grassland biomass is in the soil. The enlargement in this drawing
shows a single root with root hairs and the hyphae of mycorrhizal fungi. A film
of water (stippled area) coats each soil particle and provides habitat for nematodes and numerous protozoans, including amoebae and ciliates. Air spaces
provide habitat for mites, insect larvae, and other invertebrates. Bacteria are
extremely abundant but are too small to illustrate at this scale. Magnification
about 15 times. Based on Weaver (1968) and Hunt et al. (1987).

species have fine, light-colored hairs (pubescence)

grazed heavily by thirsty herbivores were it not for spe-

on the leaves. Small size reduces the area exposed for

cial defense mechanisms, such as sharp spines.

heat absorption and transpiration on warm days, and

Animals, too, must be adapted for surviving in the

the light color reflects some solar radiation that might

grassland environment (table 6.1). With most of the

otherwise cause leaf warming and increased water loss.

plant biomass belowground, there are many subterra-

Most leaves of grassland plants are facultatively decidu-

nean herbivores, such as nematodes, mites, insect lar-

ous, that is, they fall or become senescent after a period

vae, and various burrowing mammals (such as pocket

when conditions for photosynthesis have not been

gophers).9 Aboveground, the primary herbivores include

favorable.

grasshoppers, ground squirrels, prairie dogs, jackrabbits,

Although grassland plants are adapted to tolerate low

and the larger mammals already mentioned. One of

water availability, some avoid drought stress by complet-

the major problems for large herbivores is obtaining

ing their growth early in the spring while water is readily

enough protein from coarse grassland plants with tough

available. Examples of drought-evading plants are spring

cell walls and high concentrations of lignin and silica.

wildflowers, such as sand lily and prairie smoke. Other

Chewing this kind of food causes rapid tooth abrasion.

grassland plants store water in tissues that remain suc-

Notably, large mammals common in grasslands have

culent throughout the dry summer, for example, prickly

teeth that continually grow from their crown. This

pear cactus and stonecrop. Succulent plants would be

adaptation began to appear in the fossil record at about

Grasslands
Fig. 6.4. Grassland plants are tolerant of grazing by large herbivores,
especially if most of the forage is consumed later in the growing season.
Nearly all the plants are perennials,
and most biomass is belowground.
This rangeland in the Laramie Basin
is in excellent condition. Problems
are inevitable near water sources and
where too many animals are confined by fences for too long in areas
that are too small.

the same time that grasslands became widespread, some


20 million years ago (during the Miocene).
Another adaptation of some large grassland mammals is the ruminant digestive system of bison, deer,
elk, and pronghorn (also shared by cattle and sheep).
The four chambers of their stomachs contain bacteria
and protozoans that facilitate digestion and the extraction of protein from coarse plant material. Regurgitation and additional chewing is an important feature of
this digestive system. Nonruminants, such as horses,
can also survive in grasslands, but they must consume
considerably more food, because it passes through their
digestive systems more quickly and with less energy and
nutrients extracted. To facilitate the digestion of plant
tissues in general, all mammalian herbivores, regardless
of their habitats or digestive systems, have longer intestines per unit of body weight than do carnivores.

Partitioning Grassland Resources


Many species of plants coexist in grasslands and occupy
nearly the same space, and many grassland animals
seem to share the same habitat. However, careful study
has demonstrated that there are subtle differences in
the ways that different species obtain what they need
to survive. Some plants, for example, complete their
growth at a different time of year than their neighbors.
Plants and animals that use different resources in the
same community or at different times of the year are
viewed as occupying different ecological niches. There
is a tendency during evolution for the niches of coexisting species to become less similar, reducing the intensity of competition.
In the 1950s, John Weaver at the University of
Nebraska conducted one of the first studies that suggested resource partitioning among grassland plants. He

87

88

Plains and Intermountain Basins

Table 6.1. Some characteristic mammals, birds, and reptiles found in grasslands, shrublands, and escarpments at
low elevations

rooted plants can also benefit from water that percolates


to greater depths after snowmelt or heavy spring rains.11
While studying grassland root systems, Weaver would
sometimes dig by hand to a depth of 10 feet or more.

MAMMALS
Badger

Ground squirrel, Wyoming

Bobcat

Jackrabbit, black-tailedb

Chipmunk, least

Jackrabbit, white-tailedc

Cottontail, desert

Mouse, deer

Cottontail, Nuttalls
Coyote

Mouse, northern
grasshopper

Deer, mule

Pocket gopher, northern

Fox, red

Prairie dog, black-tailed

Fox, swift

Prairie dog, white-tailed

Ferret, black-footeda

Pronghorn

Ground squirrel,
thirteen-lined

Woodrat, bushy-tailed

On at least once occasion he decided it was prudent to


abandon the trench because of caving soil. In a sense,
the grassland is the reverse of a forest, where vertical
stratification exists aboveground, with trees above the
shrubs and herbaceous plants in the understory. Forest ecologists must watch for falling branches and even

b
c

whole trees, whereas grassland ecologists think more


about being buried in their trenches.
Grassland plants also partition resources seasonally, with cool-season plants completing their growth
by early summer or in the fall (when moisture is available), and warm-season plants growing in the summer
(until moisture becomes limiting). Junegrass, threadleaf
sedge, and western wheatgrass are examples of cool-

BIRDS
Bluebird, mountain

Meadowlark, western

season species; blue grama, buffalo grass, and little

Bunting, lark

Nighthawk, common

bluestem are examples of warm-season plants.12 This

Eagle, golden

Owl, burrowing

differentiation is based on their physiological charac-

Falcon, prairie

Sparrow, Brewers

teristics. Cool-season species commonly have lower

Grouse, sage

Sparrow, lark

optimal temperatures for photosynthesis and use C3

Harrier, northern

Sparrow, sage

metabolism. Such plants are usually less tolerant of high

Hawk, red-tailed

Sparrow, vesper

temperatures, thus growing early in the spring before

Jay, pinyon

Thrasher, sage

water stress develops. In contrast, warm-season species

Kestrel, American

Towhee, green-tailed

have a higher optimal temperature for photosynthesis,

Lark, horned

Vulture, turkey

can tolerate higher light intensities, and have a higher


water-use efficiency (that is, more photosynthesis per

Longspur, McCowns

unit of water uptake)all characteristics of C4 metabo-

REPTILES

lism. The separation of grasses into C3 and C4 species

Bullsnake

Lizard, northern sagebrush

has stimulated great interest among biologists and land

Rattlesnake, prairie

Lizard, spiny

managers, but it is only one of many ways that grass-

Now rare.

Eastern grasslands of Wyoming.

Western two-thirds of Wyoming.

land plants are adapted to their environment.


Such differences in physiology also appear to affect
responses to grazing and geographic distribution. With
regard to grazing, a critical time for ranchers and their

found that the root systems of some plants are mostly

livestock is in the spring, when hay supplies may be low

deep, whereas others are mostly shallow. In this way,

and there is little new forage production on rangelands.

neighbor
ing plants tap different parts of the soil for

The first plants to become green are the cool-season C 3

water and nutrients. To illustrate, 85 percent of blue

speciesthe plants most likely to be grazed first. The

grama roots are in the top 8 inches of soil. Neighbor-

warm-season species become important later, but by

ing plants with deep root systems, down to 3 feet or

that time there is more food available and less intense

more, include scarlet globemallow, skeleton plant, and

grazing pressure on any one group. Consequently,

slimflower scurfpea. All plants depend to some extent

repetitive early-spring grazing commonly leads to a

on surface soil derived from summer rain, but deeper-

decline in C3 species and an increase in some warm-

10

Grasslands

season species, probably because of less competition

The solar energy used for photosynthesis in grass-

from the plants that grew earlier and were eaten.13 With

lands, and all ecosystems for that matter, is typically

regard to geographic distribution, research has shown

less than 2 percent of the total impinging on the land-

that C3 species are more common at higher elevations,

scape. Nearly all remaining solar energy is reflected,

where one would expect cooler temperatures and less

used for the evaporation of water, or for heating the soil

water stress, and C4 species tend to be more common

and other ecosystem components. The low percentage

at lower elevations, where it is warmer and more often

of the solar radiation fixed by photosynthesis leads to

drier.14

an annual total net primary production (above- and

Coexisting animals also illustrate resource partition-

belowground) of 500900 g/m 2 in mixed-grass prairie,

ing. For example, bison, jackrabbits, and some grass

enough to maintain a diverse fauna of herbivores and

hoppers tend to eat more grasses, whereas pronghorn

carnivoreslarge and small.17

and some insects (including other grasshopper species)

As expected in semi-arid ecosystems, plant growth

eat more forbs.15 Thus, competition between representa-

in grasslands is strongly influenced by water availability

tives of these two groups is minimized. When the food

(fig. 6.5), though not in a linear manner. As the amount

habits of bison, cattle, pronghorn, and sheep are com-

of water increases, nitrogen can become limiting.18

pared, bison and pronghorn are most different, cattle

Notably, plant growth also is stimulated indirectly by

are similar to bison, and sheep are similar to pronghorn.

a heavy rainfall event that facilitates a pulse of organic

From an evolutionary perspective, bison and pronghorn

matter mineralization, thereby providing more nutri-

would be expected to have the most dis


similar food

ents. Various disturbances also can influence growth, as

preferences because they have coexisted for the longest

described later in this chapter.19

time. Similarly, elk and deer have coexisted for millions

Two other factors that influence plant growth in

of years, and their food preferences are quite different.

grasslands are the rate of water infiltration into the

Elk consume more grass, and deer more twigs and leaves

soil and whether precipitation comes as snow or rain.

of broad-leaved plants (called browse). Of course, the

Rain that does not infiltrate the soil evaporates quickly

diets of all coexisting animals are determined to some

and is unavailable to plants. Therefore, because coarse-

extent by what is available. Considerable overlap may

textured soils have higher infiltration rates, they often

occur in some seasons.

have higher rates of plant growth (the inverse texture


effect; see chapter 3). If more of the precipitation comes
as snow, the meltwater becomes available more slowly,

Grasslands from an Ecosystem Perspective

and there is more time for infiltration before the water


evaporates or is transpired. Generally, the mixed-grass

Energy Flow

prairie receives the majority of its annual precipitation

As in all terrestrial ecosystems, green plants make up

as rain, though snow can accumulate on the leeward

more of the grassland biomass than any other group of

sides of ridges and some taller plants (such as shrubs).

16

Plant growth and other aspects of grassland activity

The next largest component is not the large herbivores,

are often episodic, with pulses of growth, decomposi-

as is often assumed, but rather microscopic organisms,

tion, and nutrient transfers occurring after rain wets the

such as bacteria, fungi, mites, nematodes, and proto

soil. Moreover, because the rainfall is often patchy, the

zoans. Of these organisms, the weight of fungi has been

amount of plant growth and available biomass varies

estimated at four times that of bacteria. Earthworms are

from place to place. In African grasslands and savan-

rare in semi-arid grasslands, but there may be several

nas, patchy rainfall patterns determine the movements

million nematodes and a hundred thousand mites in

of large herbivores and the nomadic people associated

one cubic yard of grassland soil. Surprisingly, the total

with them. Patchy rainfall could also have affected the

weight of all mammals, birds, and insects combined is

movement of people, bison, and other animals in west-

less than 1 percent of the ecosystems living biomass,

ern North America, distributing the grazing pressure

even in grasslands grazed by domestic cattle.

over a large area.20

organisms96 percent, according to some estimates.

89

Plains and Intermountain Basins


Fig. 6.5.The annual aboveground net
primary production (NPP) of grassland
ecosystems is correlated with evapotranspiration. NPP is the amount of new
plant biomass produced in a given area
over a known period of time; evapotranspiration is the sum of evaporation
and transpiration and integrates the
effects of temperature and moisture
availability. Each point on this graph is
for a different grassland. The response
of aboveground NPP to changes in
evapotranspiration at a single site would
be somewhat different from the response
illustrated in this graph. Adapted from
Webb et al. (1978).

500

ABOVEGROUND NPP (g/m2/yr)

90

400

300

200

100

100

200

300

400

500

600

EVAPOTRANSPIRATION (mm/yr)

700

800

By what pathways does the solar energy fixed by

zoans, mites, insects, nematodes, and small mammals

photosynthesis flow through the prairie ecosystem? For

that are mostly in the soil. Daniel Milchunas, William

green plants, 6585 percent of this energy is moved

Lauenroth, and their colleagues estimated that, of the

internally from the leaves to roots and rhizomes, and

plant growth that is consumed by herbivores annually,

the plants use about 3040 percent for their own main-

roughly one-third is eaten by cattle, one-third by insects

tenance. A large portion of the remainder is converted

(above- and belowground), and one-third by nematodes,

to stems and leaves that become what is known as stand-

with jackrabbits, rodents, and birds each consuming less

ing dead or detrital biomass at the end of the growing

than 1 percent.

season. Belowground detritus, such as dead roots and

It is interesting to compare the amount of energy

dead nematodes, adds organic matter to the soil. It also

available for human use in a plant-cow-human food

provides the base for an extensive microbial food web,

chain to a plant-human system. In general, people

enables increased infiltration rates, and provides the

obtain more food per acre if plants are consumed instead

nutrients required by plants as the detritis decomposes

of animals, because much of the plant-derived energy

(see fig. 1.6). Of course, fires can burn the aboveground

of the animals has been used for their maintenance or

biomass as well, converting a substantial portion of the

is converted to animal tissues or waste products that

energy to heat and having other significant effects in

people do not eat. Further, the second law of thermo-

some grasslands.21 During some years grasshoppers con-

dynamics dictates that every conversion of energy from

sume a large portion of the aboveground biomass.

one form to anotherfor example, from plant to animal

On average, herbivores consume only 1030 percent

tissue or from animal fat to animal movementleads

of the total annual plant growth (above- and below-

to the production of heat. This principle accounts for

ground), with more than twice as much energy going

the fact that energy flows rather than cycles through

to nematodes and arthropods in the soil than to large

ecosystems. All solar energy fixed by photosynthesis is

herbivores, such as cattle (table 6.2). Not surprisingly,

eventually converted to heat, though it may persist as

some ecologists have concluded that light to moderate

biomass, soil organic matter, or fossil fuels for centuries

levels of grazing have little or no effect on plant species

or more.

composition.22 More than plants and livestock, the food

Human food production is usually greater per unit

web of rangelands is composed of bacteria, fungi, proto

area when plants are eaten rather than meat, but it is also

Table 6.2.Estimated energy flow through a prairie ecosystem on the western Great Plainsa

Ecosystem component

Energy input

Lost by respiration

Tissue production

Production
consumption

SOLAR INPUT
Global radiation

4,155,000

Photosynthetically
active radiation

1,966,000

PRIMARY PRODUCTION
Grossb

21, 882

7,439

Netc
Aboveground

2,163

Belowground

12,280

Subtotal

14,443

HETEROTROPHS
Aboveground
Herbivores
Mammalsd
Macroarthropods

Carnivoresf

105

92

13

0.13

34

23

11

0.32

8.3

Subtotal

7.5

0.8

147

123

25

127

66

61

0.10

Underground
Herbivores
Macroarthropodsg
Nematodes

0.48

50

42

7.9

0.16

20

15

4.6

0.23

12,560

9,632

2,929

72

61

12

Carnivoresh
Detritivores
Microorganismsi
Nematodes
Others
Subtotal
TOTAL

9.2

6.7

2.5

12,838

9,822

3,016

12,986

9,945

3,041

0.23
0.16
0.27

Source: Adapted from Coupland and Van Dyne (1979).


The units are kilojoules per square meter.

Gross primary production (GPP), or total photosynthesis (g/m 2/yr), varies greatly from year to year, depending on water avail-

ability (see fig. 6.4).


c

Net primary production is calculated as GPP less the amount of GPP used by plants for their own maintenance (respiration).

Mammalian herbivores include cattle, pronghorn, rabbits, and ground squirrels; cattle grazing in the study area was light (one

yearling steer or heifer per 2530 acres for 180 days each year).
e
f

Mostly grasshoppers and other insects.

Aboveground carnivores include the coyote, fox, birds of prey, and snakes.

Mostly mites and insect larvae.

h
i

Belowground carnivores include mites.

Bacteria, fungi, and Actinobacteria.

92

Plains and Intermountain Basins

true that the only way humans derive food from semi-

ing infiltration and the rate of plant growth. Fire, soil

arid rangelands is from livestock raised for meat and

disturbances, and grazing (above- and belowground) can

milkaside from hunting native animals and gather

stimulate plant growth as well, because they increase the

ing edible native plants. The alternatives are dryland or

rate of mineralization and improve nutrient availabil-

irrigated agriculture, which may lead to soil erosion and

ity.23 A simple mechanism by which mammals contrib-

other environmental problems.

ute to this process is through the deposition of urine and


feces, both of which are easily mineralized.

Nutrient Cycling in Grasslands

Nutrient loss through leaching is an unlikely phenomenon in most western grasslands, because, unlike

For their survival, organisms require nutrients as well as

in forests, rarely is there enough precipitation to cause

suitable temperatures and adequate energy and water.

water percolation below the rooting zone. Wind erosion

For example, calcium is required for plant cell walls;

during droughts or episodes of heavy grazing can lead

phosphorus is important for the storage and release

to some nutrient loss, but only a very small proportion

of energy during metabolism; potassium is needed for

of the nutrients in the soil are lost in this way. Nitro-

the regulation of cell water and manganese for the syn

gen can be lost during inevitable grassland fires, but

thesis of chlorophyll; and nitrogen is an important

again, most of the nitrogen remains unaltered in the

component of amino acids, proteins, and chlorophyll.

soil. Some volatilization of nitrogen has been reported

All nutrients cycle through grassland ecosystems, with

from animal urine and feces, but overall, nitrogen tends

losses occurring primarily during erosion events. Those

to accumulate in western grasslands (fig. 6.6). Nitrogen

losses that do occur are usually offset by nutrient inputs

loss attributable to denitrification bacteria is very small

from rock weathering, rain, snow, the deposition of dust

in mixed-grass prairie, if it occurs at all.

and finer aerosols, and such nutrient-specific processes

Because

nutrients

accumulate,

grassland

soils

as nitrogen fixation. Studies on cycling must consider

can become quite fertile and are excellent for crop

the rates at which nutrients are being added and lost,

productionassuming water is available (by irriga-

as well as the rates and pathways by which they move

tion or otherwise) and temperatures are suitable for

through the ecosystem. All aspects of cycling in grass-

the desired crop. Of course, the harvesting of crops is

lands cannot be considered here, but a few processes

itself a nutrient drain; fertilization is required after a

seem particularly interesting and relevant when think-

few years of cultivation.

ing about the western Great Plains.

Nitrogen fixation by bacteria in the nodules of

As noted, grassland plants typically have leaves and

legumes and other plants is an important long-term

stems high in lignin and cellulose, which are resistant to

nitrogen source for many ecosystems. In shortgrass

decomposition. For this reasonand because the warm,

and mixed-grass prairie, such legumes as locoweed,

moist conditions required for decomposition last for

lupine, milkvetch, scurfpea, and vetch are known to

only a short period each yeardead plant material tends

have nitrogen-fixing bacteria, and there is evidence

to accumulate on the soil surface as well as in the soil.

that a few nonleguminous plants may have the same.24

Among other effects, this surface litter (often referred

Lichens have been identified as nitrogen fixers in some

to as mulch or detritus) can improve infiltration rates.

desert shrublands,25 but so far there is no evidence that

In addition, soil organic matter increases the water and

its done by the most common lichen in Wyoming

nutrient storage capacity of the soil while providing a

grasslandsXanthoparmelia chlorochroa (fig. 6.7).

more erosion-resistant soil structure. Soil organic mat-

Symbiotic nitrogen fixation is commonly thought to

ter is resistant to decomposition, but nematodes, mites,

be the single most important mechanism that provides

and other microbes are able to use it as a food source,

nitrogen for plant growth, in the form of ammonium

directly or indirectly, and in the process nutrients from

and nitrate, but apparently this is not true in western

the organic matter are made available for plant growth.

grasslands, where legume density is low and suitable

Dead microbes also become part of soil organic matter.

environmental conditions for nitrogen fixation occur

In general, many processes create a friable soil, enhanc-

for only a short time each year. Nitrogen fixation by

Grasslands
Fig. 6.6.The amount of nitrogen, measured as grams per square meter, in the
major components (boxes) of a grassland
ecosystem, and the flows between the
components (arrows), measured in grams
per square meter per year. Based on data
in Woodmansee et al. (1981).

NITROGEN

Volatilization
Animals
0.01

Stems
Lvs
3
Roots
7

Soil
solution
2

2.3

24

Detritus
6

1.3

Animals
0.12

Atmospheric
deposition Fixation
0.5
<0.1

Soil surface

19
Soil organic
matter N
333

Microbes
5.2

Outflow (erosion)
0.1

bacteria accounts for only about 7 percent of the annual

This observation has led some managers to conclude

nitrogen inputs on a short-grass prairie south of Chey-

that fertilizing native rangelands might be profitable.

enne. 26 More important sources are rain, snow, dust,

Indeed, more forage of higher protein content could

and the microbial mineralization of soil organic mat-

result. However, several concerns have been identified:

ter, of which there is a large quantity. The sources are

(1) fertilizing only part of a rangeland could result in

diverse, but evidence suggests that nitrogen is a limiting

selective overgrazing as livestock seek out more nutri-

factor for plant growth in grasslands, especially during

tious forage; (2) certain undesirable weeds may be

wetter-than-average growing seasons.27

favored; and (3) despite increases in forage quality and


quantity, fertilization is not yet economical. Changes in

Fig. 6.7. Xanthoparmelia chlorochroa, a common lichen on the


soil surface in Wyoming grasslands and shrublands.

species composition following fertilization are typically


small unless the amounts applied are massive (>150
pounds of nitrogen/acre/year), but there is a chance that
cool-season species could be favored, because they grow
in the spring when nitrogen is more likely to be limiting
than water.28
Although annual nitrogen inputs may be low, this is
of little consequence if losses are also low and if there
is considerable nitrogen in the belowground biomass
and soil organic matter. Francis Clark, an agricultural
research scientist in Colorado, concluded that most
plant needs for nitrogen aboveground are met by the
transfer of stored nitrogen from the root system to the
leaves, with additional nitrogen being made available
by the mineralization of organic matter and microbial

93

94

Plains and Intermountain Basins

tissue, plus atmospheric additions. In 1977 Clark wrote,

stored in roots is adequate for new growth (regrowth).

Once a given N atom makes its initial entry into the

In such cases, there is hardly any disturbance at all. 31

blue grama plant, there is a greatly increased probabil-

The effect of natural grassland disturbances is to

ity that the atom will again enter new herbage growth

augment an already high level of spatial variability. As

in each of several following years.29 More than half

described by ecologist Samuel McNaughton in 1985,

(5080 percent) of the nitrogen in the senescing leaves

while writing about African grasslands, there are con-

of western wheatgrass and blue grama is transferred to

tinual fluctuations of rainfall, grazing, nutrient availabil-

belowground perennial plant parts, thereby conserving

ity, and fire. These fluctuations generate pulses of rapid

the nutrient for use when growth resumes.

plant growth in different parts of the landscape, causing

As with plant growth, the accumulation and cycling

a nomadic way of life for large herbivores and humans

of nutrients is not uniform across the grassland land-

alike. Moreover, because of this nomadism, plants in

scape. Many years of natural erosion from ridgetops,

any particular place are not grazed continuously. Today,

with accumulation on lower slopes and in valley bot-

in North America, comparable animal movements over

tomscombined with pronounced differences in the

large areas are often impeded by exurban developments

microclimate and moisture availabilityhave led to

and various land management practices.

more productive soils where sediments are deposited.30


All aspects of nutrient cycling and the flow of energy
and water vary from one place to another. The effect

The Effects of Grazing

of this patchiness on plant and animal abundance is a

As noted, herbivores of any kind form a small portion

major topic in modern ecological research.

of the ecosystem biomass, but they play an important


role in determining the nature of grassland ecosystems.

Disturbance and Succession in Grasslands

Studying the effects of herbivory provides insights on


how coexisting plants and animals survive.

A common theme in ecological research is secondary suc-

On the surface it appears as though any grazing

cession, that is, the changes that occur as an ecosystem

would be disadvantageous to plants; aboveground, the

recovers after a disturbance. All ecosystems are sub-

leaf area for photosynthesis is reduced, and energy that

jected to various kinds of disturbances, whichthough

might be expended for the production of additional

often viewed as unfortunate by casual observershave

leaves is diverted to replacing stems and leaves that have

occurred for millennia and allow for the coexistence of

been eaten. However, some grassland plants seem well

more plant and animal species than otherwise would

adapted to tolerate considerable grazing pressure, and

happen. Also, nutrient cycling and plant growth are

ecologists have hypothesized that grasslands subjected

often more rapid after a disturbance. As described

to light to moderate grazing have higher plant cover,

later in this chapter, the suppression of disturbances to

more growth, and more species than do comparable

which some native species are adapted can lead to unde-

grasslands where herbivory is excluded (fig. 6.8). 32 With

sirable consequences.

a few exceptions, however, such studies have focused

Potential disturbances in grassland ecosystems,

on mammalian herbivores rather than the micro-

aside from plowing for crop production, are drought,

scopic belowground grazers that account for most

fire, periodic heavy grazing by native or domestic her-

energy flow. Scientists can construct fences to exclude

bivores, and the burrowing of small mammals and the

large mammals, but excluding nematodes, mites, and

western harvester ant. Each disturbance has the poten-

other belowground herbivores is necessary to test the

tial to modify the relative abundance of the different

hypothesisa formidable challenge. Not surprisingly,

species. Some disturbances, such as fire or drought, are

experiments on the effects of mammalian herbivory

largely physical phenomena, whereas herbivory and

alone have produced mixed results.

burrowing are biotic. Recovery to predisturbance condi-

How might mammalian herbivory lead to increased

tions often occurs in a few months or a year if damage is

plant growth in grasslands? There are several possible

restricted to aboveground plant parts and if the energy

mechanisms:

Grasslands

growth is increased to some optimal level by light to moderate grazing (see fig. 6.8), after which continued grazing

EFFECT ON NPP

pressure leads to a decline in growth because of an excessive drain on the energy stored in the root system and less
efficient water use (caused by soil compaction that slows
the rate of infiltration). Experiments on western wheat-

Grazing intensity

grass found that, after two days of clipping, the rate of

than for plants without clipped leaves.35 The increased

photosynthesis per unit leaf area was 510 percent higher


rates of photosynthesis did not completely compensate
for the photosynthesis that would have occurred had the

Fig. 6.8.Three hypothesized relationships between large


ungulate grazing intensity and plant growth (expressed as net
primary production; NPP). Curve A, the overcompensation
hypothesis, suggests that grazing stimulates NPP up to some
threshold level, after which the effect is negative; B illustrates
the partial compensation hypothesis, which states that plants
can maintain NPP despite some grazing on some plants; and
C suggests that any amount of grazing by large herbivores
causes a decline in NPP. Adapted from Heitschmidt (1990).

total leaf area remained on the plant, but partial compensation is suggested by the results.
Several investigators, most notably ecologists Elizabeth Painter and Joy Belsky, concluded in 1993 that few
data support the hypothesis that grazing benefits plants,
but the idea has appeal and will continue to guide future
experiments. Grassland ecologist Samuel McNaughton,
one of the early proponents of the compensatory growth
hypothesis, observed that compensatory growth did
not completely replace the vegetation consumed by herbivores and that it is improper to conclude that grazing
is strictly advantageous to the plants.36

1. If some older or senescent leaves and stems are


removed by the herbivore, more light is available for
photosynthesis in younger leaves. This mechanism
may be especially important on moist sites, where
the plants are comparatively tall.
2. If water is limiting at the time of grazing, less leaf
area for transpiration might increase water availability for the remaining leaves, possibly allowing for
more photosynthesis than had occurred before the
grazing started.
3. Herbivores improve nutrient availability for plants
by concentrating certain elements in feces and
urine, which are mineralized more readily than
plant detritus.
4. By eating the stems and leaves, herbivores inadvertently stimulate the plant to produce more cells and
initiate new growth. This process, known as the
removal of apical dominance, triggers the growth of
new leaves and stems that may be more active physiologically than the older plant material that was eaten.33

Grazing also can lead to changes in the abundance


of different kinds of plants. Some species are called in
creasers, because they increase in relative abundance
with grazing pressure, whereas others are classified
as decreasers, because they become less abundant. At
some point, grazing pressure, whether from native or
domestic herbivores, becomes so heavy that weedy
species (invaders) increase in abundance. Range managers have learned to use trends in the relative abundance
of increasers, decreasers, and invaders to assist with
evaluating range conditions. Today, however, managers
think more commonly about the ecosystem services
provided by a grassland, even if the species composition
is different than what might have been considered ideal
40 years ago.
Both bison and cattle undoubtedly grazed western
grasslands heavily at times. 37 Explorers passing through
Fort Laramie observed what they thought was overgrazed rangeland, and one noted that the area was
barren country abounding with prickly pearsa cactus still common today. Another traveler at about the

With such mechanisms in mind, the compensatory

same time reported that 125,000 head of cattle had

growth hypothesis was proposed.34 It states that plant

been driven through the same area, leaving the land

95

96

Plains and Intermountain Basins

as devoid of grass as the streets of New York.38 Pre-

quickly throws the advantage in competition to the

sumably the cause was livestock abundance, perhaps in

side of the latter. Because of more water and light,

combination with drought. A hundred years later, the

their growth is greatly increased. They are enabled to

grasslands of this area appeared to have recovered. The

store more food in their propagative organs as well as

rate of recovery may be slow or rapid, depending on the

to produce more seed. The grazed species are corre-

plant species present at the time, weather conditions,

spondingly handicapped in all these respects by the

the history and nature of grazing pressure on the site,

increase of the less palatable species, and the grasses

and livestock management during the time of recovery.

are further weakened by trampling as stock wan-

Some plant ecologists attribute pricklypear cactus

der about in search of food. Soon bare spots appear

abundance to excessive livestock grazing, but Lewis and

that are colonized by weeds or weed-like species.

Clark observed in 1805long before the first domestic

The weeds reproduce vigorously and sooner or later

livestockthe prickly pears are so abundant that we

come to occupy most of the space between the frag-

could scarcely find room to lye [sic].39 Various investi-

ments of the original vegetation. Before this condi-

gators have observed how cactus increases with drought

tion is reached, usually the stock are forced to eat

and decreases following wetter than normal years, pos-

the less palatable species, and these begin to yield to

sibly because of greater susceptibility to insects. There

the competition of annuals. If grazing is sufficiently

is little or no evidence that cactus biomass increases

severe, these, too, may disappear unless they are

because of grazing pressure alone.40

woody, wholly unpalatable, or protected by spines.43

Travelers in Wyoming often wonder whether the


grasslands and shrublands are similar to those observed
prior to the arrival of Europeans. The photographic
record suggests that only a few changes have occurred,
but it is not known whether all the plant species are still
present. Were some plant species driven to extinction
by livestock grazing during the 1880s? Such questions

Lincoln Ellison, who in 1960 provided one of the


early comprehensive reviews of the effects of grazing on
grasslands, recognized that North American rangelands
have been subjected to grazing for millennia. He made
the following observation when comparing fires, plowing, and grazing:

are difficult to answer. Ecologists Richard Mack and

In arid lands, fires may be rare because the fuel is

John Thompson concluded in 1982 that Great Plains

eaten by grazing animals, denudation by plowing

plants are more tolerant of domestic livestock grazing

may be sporadic because of marginal returns and

than are plants in the Great Basin to the west. They

recurrent drought; but overgrazing, although caus-

attributed this difference to a much larger population

ing less complete denudation in any one season, is

of bison on the Great Plains for thousands of years, with

important because it is widespread year after year.44

the result that plants there evolved in the presence of


substantial grazing pressure.41 Long-term studies south
of Cheyenne suggest that changes in annual rainfall
affect plant cover and species composition far more
rapidly than livestock grazing does, and though heavy
grazing may cause declines in the abundance of some
palatable species, moderate grazing (defined as 50 percent removal of plant biomass) did not cause a decline
in range condition.42
The mechanism by which heavy grazing affects
grassland vegetation is generally well known and was
described, perhaps most succinctly, by Nebraska ecologists John Weaver and Frederick Clements in 1938:

In some areas, rangelands probably are still recovering


from the adverse effects of livestock in the late 1800s
and early 1900s, prior to the advent of modern management practices.
Notably, not all grasslands respond to grazing pressure in the same way. Daniel Milchunas, William Lauenroth, and their associates in 1988 compared grasslands
worldwide over a broad range of climatic conditions and
concluded that those in eastern Wyomingin a semiarid environment and with a long history of grazing by
such native ungulates as bisonare tolerant to grazing pressure. In contrast, the taller grasslands of more
humid climates, and grasslands in areas with no history

The more palatable species are eaten down, thus

of long-term grazing, change rather dramatically with

rendering the uneaten ones more conspicuous. This

the introduction of livestock.

Grasslands
Fig. 6.9. Mixed-grass prairie
in Wind Cave National Park,
with bison grazing in a
prairie dog town. Elevation
4,100 feet. Photo by James K.
Detling.

Of course, any grassland can be affected adversely

Serengeti appear to depend on wildebeest for their

if too many animals are fenced into small areas.45 In

survival, because the wildebeest eat the coarser for-

1991 Michael Coughenour described how the nomadic

age, thereby stimulating the growth of the new foliage

movements of large herbivores, triggered by the greater

required by gazelles.48 A month or so after the migra-

availability of preferred food nearby, constitute an im

tory wildebeest move through an area, consuming up

portant mechanism that enables some grasslands to tol-

to 85 percent of the green biomass, regrowth of the

erate grazing, simply because the animals seek food in

plants creates a grazing lawn on which the gazelles

new areas as forage becomes depleted where they are.46

feed. Samuel McNaughton, an expert on this eco

Various livestock management systems, such as short-

system, suggested that the magnificent carnivores in

duration grazing and rest-rotation grazing, are based

African savannas are more numerous because of the

on this principle, but there is still considerable debate

nutrient and energy flows facilitated by the herbivores

over their effectiveness, probably because generaliza-

and their interactions.

tions are difficult. Each ranch operation requires its

North American grasslands have had far fewer large

own management plan based on goals developed with

mammals since the great mammalian extinction about

an understanding of the environmental constraints of

12,000 years ago (see chapter 2), but interesting interac-

the available land.47 Rangeland ecologist David Briske

tions occur here also.49 Most notably, research by James

and his associates concluded that rangelands should be

Detling and his associates in Wind Cave National Park in

managed more to restore ecosystem services, and that

South Dakota have shown that bison, a nomadic herbi-

rotational grazing may not be the best approach when

vore, graze in prairie dog colonies (towns) more often

climate change, invasive species, fragmentation, and

than would be expected by chance (fig. 6.9). The forage

the conservation of biodiversity are concerns.

on the colonies is more nutritious than off, apparently

Research on African grasslands, where 25 or more

because more animal waste is deposited there. Nitrogen

species of large mammalian herbivores occur, has

uptake by the grazed plants is greater, and there are

revealed several interesting interactions. One is the

more young leaves (which tend to be relatively high in

way in which some African herbivores improve food

nitrogen content) in the available forage. In addition,

availability for others. To illustrate, gazelles on the

the regrowth of young shoots is stimulated by grazing.

97

Plains and Intermountain Basins

Similarly, Krueger found that pronghorn used prairie

50

dog colonies preferentially, perhaps because of a higher


number of forbs there.

Cause of fire in 1960:


Human

50

Lightning

Such interactions among bison, pronghorn, and


40

burrowing animals surely must have been widespread


prior to the arrival of Europeans. Because of the prairie
dog and other burrowing animals, North American
bison may have been more abundant, along with their
predators (grizzly bears and wolves). One study found
more herbivorous insects on rangeland grazed by cattle
than on rangeland that had not been grazed for a long
period, and another study found more belowground
invertebrates in the soil of prairie dog towns than away
from the towns.51

NUMBER OF FIRES

98

30

20

The effects of large herbivores have attracted the


attention of most grassland scientists, yet the effects of
the abundant belowground herbivores, small as they

10

are, should not be ignored. For example, in an experimental study, a nematocide was applied to kill all the
nematodes in the study area.52 Plant growth declined.

Apparently, microbial grazers improve nutrient availability for vascular plants just as their much larger
aboveground counterparts do.

APR MAY JUN JUL AUG SEP OCT

Fig. 6.10.Number of recorded human- and lightning-caused


fires in the grasslands and shrublands of the Powder River
Basin north of Douglas, Wyoming, in 1960. Adapted from
Komarek (1964).

Fire
Prior to the advent of fire suppression, prairie fires

Most occur during July and August, when fuels are

occurred frequently because of the highly flammable

drier and thunderstorms more frequent. Thunder-

leaves and stems that rapidly accumulate. Without fire,

storms may produce enough rain to extinguish such

and if grazing intensity is low, the standing biomass

fires, but lightning strikes with no or little rain are

is gradually flattened to become part of the mulch on

common. 55

the soil surface. Once fires were ignited, plumes of

Native Americans also started fires, sometimes

smoke would have been visible. The fires would have

to facilitate their hunting.56 In 1985 George Gruell

burned for weeks or months until extinguished by wet

reviewed 145 historical accounts of fire by 44 observers

weather, or until an effective fire break was reached

in the Rocky Mountain region. He concluded that fires

(such as a sufficiently wide river or ridge, or an area of

set by Indians were most common in the lowlands and

inadequate fuel to sustain the fire). Fires might have

that they could have been annual events in some areas

been less common during extended droughts, largely

(although probably the same tract of grassland did not

because of less fuel.

burn in two consecutive years). For much of the Great

53

The source of ignition was usually lightning strikes,

Plains, mean fire intervals were from 2 to 25 years, with

which start fires in grasslands as they do in forests.

the longer intervals in areas of rough topography.57

An annual average of 1025 lightning-caused fires

Grassland fires in central Nebraska occurred at intervals

per 1,000 square miles has been reported for differ-

of 46 years from 1850 to 1900, and every 1530 years

ent parts of the northern Great Plains. 54 Lightning has

in western Nebraska. Wyoming grasslands probably

been documented as the major cause of prairie fires

burned less frequently owing to slower fuel accumula-

in the Powder River Basin north of Douglas (fig.6.10).

tion in a more arid climate (fig. 6.11).

Grasslands
Fig. 6.11. Late-summer fire in the
mixed-grass prairie of the Laramie
Basin. Fires in tallgrass prairies are
less easily extinguished. Photo by Ken
Driese.

Generalizations about the effects of fire on specific

soil to warm up and plant growth to begin earlier in

plants are difficult, but warm-season species seem to

the spring, when water is available. Very little nitrogen

have evolved a higher level of tolerance for burning

is lost to the atmosphere by burning, compared to the

than cool-season species, probably because they are

amount that remains in grassland soils. Furthermore,

more likely to be growing in the drier summer months

the amount lost is replaced in a few years by precipita-

when most fires occurred. Some species increase in

tion (see fig. 6.6).

abundance after fires, not because fires are directly

Because of these influences, total plant growth is

beneficial, but because additional water and nutrients

sometimes higher after a fire, at least in relatively moist

become available following the death of less-tolerant

prairies with tall grasses, such as in central and east-

species, or because a more favorable environment for

ern Nebraska and in eastern Kansas. This observation

their growth is created by the burning of mulch on the

has led some managers to encourage the burning of

soil surface.58

grasslands as a regular practice. However, mixed-grass

Fires also affect nutrient cycling and plant growth.

and shortgrass prairies in the West are more frequently

In some cases plant growth following fire is faster, with

water stressed than tallgrass prairie, a condition that

greener, more palatable nutrient-rich stems and leaves.

could be aggravated by fire because of increased evapo-

Burned grasslands initially appear black and totally dev-

ration from the warmer, blackened surface. 59

astated, but in a few weeks it is common to see herds of

The effects of grazing and fire on plant growth are

bison, pronghorn, and cattle grazing on the regrowth.

interactive. Grazing can be so intense that there is little

Indeed, Indians probably burned grasslands intention-

mulch left on the soil surface, which leads to lower flam-

ally to attract game to smaller areas, where they could

mability. If, however, a large amount of mulch accumu-

be more easily harvested. The animals would have been

lates due to fire suppression or very little grazing, then

lured by more nutritious forage, resulting from the sud-

much of the water in short summer showers can be

den availability of nutrients after accumulated mulch

intercepted by the mulch and evaporated without ben-

was burned. Combustion, like microbial decomposi-

efiting the plants at all. Such adverse impacts may be

tion, converts organic matter to the inorganic nutrients

offset to some extent by the increased trapping of snow

needed for plant growth.

by the accumulated biomass and less evaporation from

Another effect of burning, especially in taller grasslands, is to remove the insulating mulch, allowing the

the shaded soil surface, but this beneficial effect is not


known to happen regularly.

99

100 Plains and Intermountain Basins

Rather than consider the effects of grazing and

and shortgrass prairie could be used for crop produc-

burning separately, ecologists now sometimes recom-

tion in a manner similar to the tallgrass prairies to the

mend a patch-burn grazing system, where the grassland is

east. The consequences were staggering dust storms. A

exposed to both livestock grazing and relatively small,

subsequent drought occurred from 1952 to 1955, but

patchy burns in the same management unit, enhanc-

by that time land management practices had improved

ing the probability of higher biodiversity because of the

and blowing dust was less severe.

more heterogeneous landscape that results. Large herbi-

After the end of an extended drought, the recovery

vores are attracted to the recently burned area because

of mixed-grass prairie may require several years or more,

of more palatable and nutritious forage there, for rea-

with some species reinvading from the shelter of prickly

sons just described. Yet nearby, some birds and small

pear cactus.64 Livestock managers do not usually appreci-

mammals thrive on the tracts of unburned grassland

ate cacti, but spiny plants enable other plants to survive.

that remain.60

Forage availability increases dramatically after cactus are


burned, partly because of greater accessibility without

Drought
The continental climate of the western Great Plains

the cactus spines, but also because plant growth is higher


without competition from the cactus. Still, pricklypear
cactus can help maintain biological diversity.

has great fluctuations in annual precipitation, and its

With time, many of the original species become re

common to have several consecutive years with well-

established following extended drought. Seedling estab-

below-average amounts of rain and snow. Predictably,

lishment is more probable with increased precipitation,

plant growth and rangeland carrying capacity are much

but new shoots also sprout from surviving rhizomes.

lower during droughts. Just as important, plant cover is

Research suggests that some perennial natives survive

reduced and shifts in species composition may occur.

droughts in a dormant or near-dormant state, even for

Some research suggests that shifts in species com-

periods of 57 years. Above-average rainfall for 13years

position commonly attributed to heavy grazing could

changed a shortgrass prairie dominated by blue grama

be the result of drought.61 Common Wyoming plants

to a mixed-grass prairie with needle-and-thread grass,

that sometimes increase in abundance during dry peri-

prairie junegrass, and western wheatgrass as well as

ods include blue grama, Sandberg bluegrass, prickly

blue grama.65 Predictably, livestock forage increased

pear cactus, and threadleaf sedge. Those that decrease

considerably.

include junegrass, little bluestem, needle-and-thread


grass, red threeawn, silver sagebrush, and western
wheatgrass. One of the most successful species during

Grasshoppers

drought is Sandberg bluegrass, but even it declines with

Drought disturbances are frequently accentuated by

extended drought.62 Species that survive or increase do

outbreaks of grasshoppersarguably the most impor-

so partially because of reduced competition from less

tant insect in western grasslands (fig. 6.12).66 The causes

drought-tolerant species.

of outbreaks during drought are not well understood,

The changes associated with drought are most likely

but some research suggests that bacterial and fungal

reduced plant growth, biomass, and cover. Plant cover,

grasshopper diseases are more common under rela-

for example, decreased from 69 to 2 percent during a

tively moist conditions, limiting grasshopper survival.

drought in eastern Colorado, and from 28 to 2 percent

It would be of great practical significance to know that

in southeastern Montana.63 Such drastic changes might

population sizes are determined by the weather, but

be partially the result of grasshopper outbreaks, which

Gordon Watts and his associates cautioned that, even

tend to coincide with drought years.

if there is a climatic correlation with grasshopper abun-

Unfortunately, the drought years of the 1930s (1933

dance, There can be no assertion as to whether it is

1939, especially 1934) followed a 50-year period of grass-

a direct action of weather on the grasshoppers physi-

land plowing, spurred on in some areas by large markets

ology, an indirect action on the food plants, a differ-

for wheat and the erroneous notion that mixed-grass

ential effect on their predators, parasites, and diseases,

Grasslands
Fig. 6.12. More than 100 species of
grasshoppers are found in Wyoming.
Most are adapted for the grasslands and
shrublands in the lowlands, but some
species occur in the alpine tundra. This
one, known as the flambellate grasshopper (Melanoplus occidentalis), is common
in grasslands. It eats mostly forbs but
also grasses, seeds, and dying or dead
arthropods. Photo by Scott Schell.

or a measure of each.67 Favorable conditions for grass

by grasshoppers when their population sizes are high

hopper population outbreaks have also been attributed

over large areas. The costs of the various control mea-

to heavy grazing, whether by bison or livestock.68 Some

sures were justified, in part, by the expectation that they

of the most interesting and complex ecological ques-

would keep the population down for several years. That

tions pertain to the population dynamics of insects.

may be true in some areas, but a recent study concluded

The amount of energy that flows to grasshoppers can

there is little basis for prorating the benefits of con-

be small or large. In northern Colorado they consume

trol beyond the year of treatment and that managing

less than 0.5 percent of the plant growth during years

grasslands in ways that conserve the natural predators

with normal population sizes, but 63 percent of the

of grasshoppers is likely to be a more sound approach,

forage is eaten in outbreak years, with about 20 grass

ecologically and economically.72

hoppers per square yard.

Notably, the apparent veg-

All Wyoming grasshoppers are native and, like bark

etation damage is not a good measure of the amount

beetles in mountain forests, are integral components of

69

actually eaten by grasshoppers, as they cut up to 25

the grassland ecosystem. Curiously, the most notorious

times more forage than they eatleaving considerable

grasshopper in the region, the Rocky Mountain locust,

detritus on the soil surface.

became extinct in the early 1900s. From time to time

In 1941 B. W. Allred described the effects of grass

this locust swarmed over hundreds of square miles,

hoppers in the Powder River Basin during the 1930s

forming virtual clouds that consumed crops and native

drought, writing that there were 50100 per square foot

plants alike. Plagues of flightless Mormon crickets do

and that they destroyed all edible vegetation.70 Much

much the same, as do other kinds of grasshoppers. How-

earlier, in 1864, General Alfred Sully wrote, the only

ever, their swarms are not as large as those of the now-

thing spoken about here [Montana] is the grasshopper.

extinct Rocky Mountain locust.73

They are awful. They actually have eaten holes in my


wagon covers and in the tarpaulins that cover my stores.
Curiously, some areas are not affected even during the

Prairie Dogs and Other Burrowing Animals

worst outbreak years, which illustrates another way that

Burrowing animals can cause prominent changes

disturbances can cause patchiness in landscapes.71

in grasslands and mountain meadows. For example,

For many years, ranchers and scientists have worked

pocket gopher mounds sometimes cover 25 percent of

toward the goal of minimizing economic losses caused

the soil surface.74 Predatory skunks and badgers search-

101

102 Plains and Intermountain Basins

ing for the gophers often amplify the effect with their
digging. This burrowing moves tons of soil every year,
helping to create a more friable soil structure that has
long-term benefits for plant growth. Pocket gopher
burrows provide habitat for the grasshopper mouse, as
well as the beetles and crickets on which the mice feed,
and though the gopher is nocturnal, some individuals
that make the mistake of coming to the surface during daylight provide food for raptors. Notably, pocket
gophers tend to occur near clumps of pricklypear cactus, possibly because the succulent cacti serve as a
source of water. The gophers are able to eat around the
spines.
The most studied burrowing mammal in grasslands
is the prairie dog (fig. 6.13; see also fig. 6.9). Two species
are found in Wyoming, the black-tailed prairie dog on
the Great Plains and the white-tailed prairie dog of the
intermountain basins. One of the most obvious effects
of prairie dogs is the extensive burrowing that characterizes their colonies (towns). As with pocket gophers, large
volumes of soil are moved, improving infiltration, hastening the incorporation of organic matter, facilitating
nutrient cycling, and increasing spatial heterogeneity.75
Generally, prairie dog burrows are dug to depths of 39
feet, deeper than pocket gophers. Along with the pocket
gopher and other burrowing mammals, prairie dogs
enable a higher level of biological diversity by creating a
variety of subterranean habitats, such as for burrowing

Fig. 6.13. Prairie dogs, along with other burrowing animals


(such as Richardsons ground squirrels, badgers, and harvester
ants), alter vegetation structure on the grasslands and shrublands of the intermountain basins and western Great Plains.
Their burrows can be 6 feet deep or more. The black-tailed
prairie dog, shown here, is found at lower elevations on the
western Great Plains; the white-tailed prairie dog is found
in the intermountain basins to the west. Photo by LuRay
Parker/Wyoming Game and Fish Department.

owls, rattlesnakes, grasshopper mice, desert cottontails,


reptiles and amphibians, which in turn encourages the

favorable. However, research suggests that large prairie

raptors that feed on such animals (most commonly the

dog populations are a symptom of heavy livestock graz-

ferruginous hawk). Large colonies also provide habitat

ing. They reason that prairie dog densities increase with

for the restoration of black-footed ferrets, which depend

heavy grazing because predator success is diminished,

on prairie dogs.76 Like the beaver in riparian zones (see

due to less plant cover around the town. Moreover,

chapter 4), prairie dogs influence grassland ecosystems

heavily grazed rangelands commonly have more forbs,

more than expected based on their low biomassthe

which are preferred by the rodents, and by pronghorn,

definition of a keystone species. Because beaver and bur-

too. The interaction of clipping, grazing, and predation

rowing mammals play such a deliberate role in creating

may be less likely in shortgrass prairie, where the veg-

their habitat, they are also known as ecosystem engineers.

etation is already short and visibility good.

In addition to extensive burrowing, the black-tailed

Although prairie dog towns add significantly to the

prairie dogs on the Great Plains keep the surrounding

biological diversity of the grasslands where they occur,

vegetation clipped close to the ground, even if they dont

they also appear to reduce the amount of livestock for-

eat itpresumably to improve their ability to detect

age. Moreover, some of the land is pockmarked with

stalking predators (see chapter 16). This clipping leaves

their burrows, creating a hazard for livestock and horse-

the impression of an overgrazed rangeland, which has

back riders alike. Consequently, like wolves, the rodents

led some to doubt that their effects on the rangeland are

are most often viewed as problem animals where live-

Grasslands

stock production is the primary management objective.

Ground squirrels are another group of small mam-

Often they are shot or poisoned even today, when their

mals that can affect vegetation through herbivory and

population sizes are a small fraction of what they were

burrowing. They do not form towns as prairie dogs do,

in the 1800s. Various studies have described how prairie

but their populations fluctuate between nearly zero and

dogs have little or no adverse effects on livestock weight

a hundred or more per acre. The causes of ground squir-

gain, and that long-term benefits for forage productivity

rel population fluctuations are still not well known.

outweigh what appear to be adverse effects. As noted,

Another common feature of the plains and basins

research has shown that the nutritive value of forage

are the mounds created by harvester ants (Pogono

on prairie dog towns is higher than in much of the rest

myrmex occidentalis).80 These seed-eating ants are some-

of the landscape, apparently because nutrient cycling

times viewed as destructive, because they denude areas

is facilitated by the activities of so many animals in a

around their dome-shaped mounds (fig. 6.14). An esti-

small area. There may be a need to control prairie dog

mated 6 million mounds existed in the Wind River

77

populations in some areas, but doing so has become

Basin in the 1970s. The common reaction to such num-

highly controversial.

bers is to lament the loss of forage for livestock. Yet, like

This issue recently came to a head on the Thun-

other burrowing animals, ants mix and aerate the soil,

der Basin National Grassland, where federal man

and the vegetation surrounding denuded areas is often

agers are trying to increase the size of a black-tailed

more vigorous, most likely because of improved water

prairie dog town so that it can sustain a population of

and nutrient availability caused by ant burrowing and

the en
dangered black-footed ferret. Appropriately, the

defecation. The absence of plant growth on the mound

managers are collaborating with adjacent landowners,

is probably compensated for by the increased growth

encouraging them to allow some prairie dog towns on

nearby (see fig. 6.14).

their property as wellat least in small colonies.78 Some

Disturbance by burrowing mammals and ants also

ranchers have consented to this plan, but only if they

adds patchiness to the grassland mosaicabove- and

can shoot or poison nearby dog towns where they are

belowground. The disturbances lead to continual

not wantedwhether on private or public land. Shoot-

change in the grassland community as species adjust

ing prairie dogs, however, does not make sense to those

to different environmental conditions. The native

who ask: How can poisoning black-tailed prairie dogs

flora and fauna are well adapted to such disturbances;

be justified when their benefits to the ecosystem are so

their diversity is high in large part because of them.

well known and they have been petitioned for protection

As in other ecosystems, however, the effect of people

under the Endangered Species Act?

has often been to modify the frequency of grassland

The dark cloud of sylvatic plaguea bacterial dis-

disturbances or to increase their intensity, resulting

ease carried by fleas spread by rodents and introduced

in a more homogeneous landscape with lower spe-

to North America with rats in the 1800scomplicates

cies diversity. In many areas, the natural disturbances

matters, because it spreads rapidly in prairie dog towns

have been replaced with new ones, like plowing over

and can cause local extinction of the rodents in a year

large areas, for which native species are not adapted.

or two. That has happened numerous times in the past

Thus, it is little wonder that introduced weedy species,

40 years.79 Insecticides have been used to kill the fleas,

originating in Eurasian landscapes with a much longer

with considerable successthough with the usual

history of agriculture, have done so well in American

expense and undesired effects of adding pesticides to

croplands.

the ecosystem. For two reasons, sylvatic plague could


foil the best-designed plan for restoring the ferret in
portions of its native habitat: the ferrets food supply

Invasive Plants

is diminished, and the ferret itself is susceptible to the

Native grasslands have been invaded in some places

disease. Vaccinating the ferrets against the disease has

by a variety of introduced plants that have predictable

shown some promise, but there are too many prairie

effects: crowding out native species, altering habitat

dogs for that to be practical.

and food availability for wildlife and livestock, reduc-

103

104 Plains and Intermountain Basins

Fig. 6.14.The absence of plant growth on harvester ant


mounds probably is compensated for by the increased growth
around the mound. Plant growth on the perimeter would

be favored by better soil aeration caused by the ants, greater


availability of water and nutrients in the affected area, and
less competition from other plants.

ing biological diversity, affecting the amount of carbon

sometimes cattle. Herbicides should be avoided when

stored in the soil and nutrients available for growth

possible, as they kill some native species as well as the

(plant and microbial), increasing soil erosion in some

target plants and affect the soil biota in unknown ways.

areas, and creating the need for expensive weed con-

The challenge seems hopeless when weedy plants domi-

trol practices. Most introduced plants do not become a

nate a large areahence the admonition by weed con-

problem, but a few outcompete native species, because

trol specialists to learn how to identify noxious plants,

they are pre-adapted to their new environment.81

monitor rangelands closely for their appearance, and

Moreover, they are free of their native pathogens and

act promptly to eliminate the plants while their popula-

species-specific herbivores. Grassland invaders on Wyo-

tions are small and isolated.

ming watch-lists include Canada thistle, tumbleweed

Many Wyoming grasslands do not yet have invasive

(Russian thistle), cheatgrass, leafy spurge, and the vari-

weeds. Most likely this is the result of management

ous knapweeds (diffuse, Russian, and spotted). Cheat-

practices that maintain the native plants in good condi-

grass has received the most attention and is discussed

tion, especially during recent decades, when the seeds

in the next chapter.

of exotic plants have become so much more widespread.

Unlike other disturbances, the effects of invasive

Essentially, the native plants are fully occupying the

plants can be long-lasting and restoration is difficult.

spaces into which exotics might invade. Another fac-

Their removal from grasslandswhere that is judged

tor may be that western grasslands are too dry for most

to be importantmay require a combination of con-

invasive plants. To illustrate, Canada thistle and leafy

trol measures, including herbicides; biological con-

spurge tend to occur in the vicinity of riparian zones

trol; hand pulling; fire; changes in land management

or irrigated land rather than in semi-arid grasslands on

practices; and targeted grazing by goats, sheep, and

the upland. Realistically, though, land managers should

Grasslands

Fig. 6.15. Less-palatable plants are more likely to invade


grasslands when plants are weakened by excessive grazing,
such as to the left of this fence. Grass cover has declined and

the abundance of Douglas rabbitbrush, a native shrub, has


increased on the heavily grazed land.

anticipate that the invasive plants of the future could be

evolve to allocate less energy to species-specific chemi-

more drought tolerant.

cal defensive compounds and more energy to growth

82

Invasive species often gain a foothold when distur-

and reproductionan option that would not be adap-

bances reduce the vigor of native species. Some of these

tive for native species.83 Some observations suggest that

are human caused, such as excessive grazing (fig. 6.15),

invasive species are not nearly as abundant during the

plowing, and road construction; but others are natural,

first few years as they are later, after they have adapted

such as drought, fire, and the burrowing of small mam-

to their new environment in one way or another.84 Thus,

mals. A modern predicament is that even natural dis-

just because an introduced plant is localized does not

turbances can create habitat for undesired plants (see

mean that it will remain innocuous. Similarly and on a

chapter 16). Their seeds are widely dispersed, and some-

more positive note, some native species after a period of

times such plants become established before native

time may evolve physiological and morphological adap-

species can.

tations for competing with invasive species.85

An increased understanding of weed ecology and the

Another insight was proposed by Agricultural Re

characteristics that determine the susceptibility of some

search Service scientist Dana Blumenthal and his col-

ecosystems to invasion has led to several noteworthy

laborators.86 They observed that invasive species were

insights. First, once established, invasive plants (and

most successful when slow-growing native plants were

other invasive organisms) adapt further to their envi-

disturbed and when native enemies of the invasives did

ronment, enabling them to become even more aggres-

not exist. The first step to reduce the spread of invasive

sive. This could involve acclimatization as well as a

plants, they reasoned, was to enable the persistence of

genetic response. For example, once in an environment

the native plants, and if possible, find a herbivore that

without parasites and pests, the invasive species may

would feed on the invaders. It is tempting to infer that

105

106 Plains and Intermountain Basins

irrigating or fertilizing the native plants would be good,

tion until a small population was found near Meeteetse

to increase their vigor, but Blumenthal and his associ-

in 1981. Now, through captive breeding, there are sev-

ates found that maintaining water and nutrients at low

eral thousand ferrets, and some are breeding success-

levels reduced the invasibility of the grassland they

fully in the wild.88

studied in northern Colorado.

Careful manipulation

The most conspicuous grazing in the twenty-first

of the time and nature of livestock grazing also would

century is by cattle. But, of the large herbivores, only

restore some level of herbivory, assuming the invasive

the pronghorn now completes its life cycle on the

species were eaten.

grasslandseach animal very likely dying in or near the

87

Where weeds are already well established and the

same county where it was born. In the early 1700s, Native

goal is to restore native species, Blumenthal and his

American hunters would have seen a greater diversity of

associates found that adding carbon to the soil, in the

large mammals, and very likely they would have heard

form of sawdust, reduced the availability of nitrogen.

the barking of prairie dogs almost every day during the

This causes a decline in soil fertility, which the native

summer. Raptors and other predators would have been

plants can tolerate better than the invasive species. In

more common with such a large prey base. There would

their experiment, adding nitrogen favored the intro-

have been many large animal carcasses, providing food

duced species. Thus, disturbances that make nutrients

for numerous kinds of scavengers. Nathaniel Wyeth

more available should be minimized while this kind

wrote in 1832 about an abundance of buffalo along the

of restoration work is in progress. Where feasible, her-

Sweetwater River valley, and in 1857long before the

bicides were integrated with this work in a way that

introduction of sheep and cattleWilliam A. Carter

favored the native species, helping them become estab-

wrote about seeing thousands of dead animals.89 Ravens,

lished. This research was motivated by the observation

vultures, coyotes, and other scavengers would have fed

that the primary expense of controlling invasives, by

on the carcasses. Now and then a grizzly bear or pack

whatever means, results from the fact that it has to

of wolves would have chased them away. Seeing a bad-

be done time and time again. If native species can be

ger would be common. The clickity-clack of grasshopper

restored and managed properly, they persist year after

wings would have been frequent in the summer, but with

year, preventing or slowing plant invasion well into the

no spotlights, the nocturnal swift fox and black-footed

futurean ecosystem service powered by solar energy.

ferret would have gone unnoticed most of the time.

Native plant growth and biomass may be less than that

Today, a grassland ecosystem like the one first

of the invasive plants, but the expense of maintain-

encountered by EuroAmericans in the early 1800s, and

ing the grassland is greatly diminished. Moreover, the

the first human immigrants some 15,000 years earlier,

adverse effects of the invading species are reduced.

does not exist. Most of the native plants are there, and
some of the small native animals can be found if one

Challenges for the Future

looks long enough. Grasshoppers still exist, some years


more than others, and occasionally a prairie dog town

Traveling across North America, its rare to find native

can be found. Coyotes will be heard. With luck, a swift

grasslands that extend to the horizon as they once did.

fox will be visible as it darts across the road at night.

Most have been plowed for cropland or converted to

However, off in the distance, the large animals are most

pastures of introduced plants. Prairie dog towns can

likely cattle rather than bison, deer, elk, pronghorn, or

be seen along the highways in the western states, but

large predatorsor the camels, mammoths, saber-tooth

the population of these animals is probably less than

tigers, and other extinct species of earlier times.

10percent of what it was in the 1800s. Many think that

Considering the relatively depauperate nature of

the black-tailed prairie dog should be listed federally

grasslands nowat least abovegroundthere is reason

as an endangered species, especially now that sylvatic

to promote the establishment of a grassland national

plague has been introduced to America and frequently

park in an area large enough to restore the diversity that

kills entire colonies. Because of the decline in prairie

once existed there. Neighbors would justifiably object

dogs, the black-footed ferret was on the brink of extinc-

to the re-establishment of wolves and grizzly bears on

Grasslands

the Great Plains, but all the other species could be re

Unfortunately, those circumstances are chang-

introduced. All dead animals would remain for scaven-

ing, leading to concerns that the values of grasslands

gers and nutrient cycling. The prairie dog towns could

will be lost. Plowing continues, often in concert with

grow to be large enough to sustain the black-footed fer-

shifts in crop prices and by landowners desiring to take

ret, assuming that the sylvatic plague can be controlled.

advantage of government subsidies. The cultivation of

Fires would be started or allowed to burn as judged

grassland soils is now more often sustainable than pre-

appropriate, and the area would be large enough that

viously, thanks to minimum till or no-till agriculture,

unintentional wildfires could burn without burning

but the grassland ecosystem is thoroughly disrupted in

the whole park. Protection for the neighbors would be

the process. Some grasslands have been restored, either

required. Some bison would be harvested to simulate the

by abandoning old fields and allowing secondary suc-

effects of the wolves and grizzly bears that are absent.

cession to occur, or through the planting of perennial

Traveling into this imagined park from any direction,

native species, such as farmers and ranchers are able to

the aboveground natural diversity of native grasslands

do with funding from the U.S. Department of Agricul-

would be immediately apparent.

tures Conservation Reserve Program. Wildlife invari-

Fortunately, some prairies in the semi-arid West

ably benefits, but decades if not centuries are required

have not been plowed, and they have been managed

before the original biological diversity is restoredif

for livestock production in a way that appears to be

thats possible at all.94 Such concerns cannot preclude

sustainable. The plants are mostly native species, the

farmers from making a living from their land, but the

all-important belowground food webs probably are still

best soils most likely have already been cultivated. With

intact, and the natural diversity aboveground is high if

so much of the grasslands already disrupted, there is

not complete. Prairie dogs and their associated species

reason to be concerned about further losses.

are present.90 Some intact grasslands are private; others

The intact grassland ecosystem requires an area that

are on public land, such as in Wind Cave National Park

is better thought of as a diverse landscape with ridges

and Badlands National Park in western South Dakota,

and riparian zones here and therenot a homogeneous

and a place known simply as Grassland National Park in

tract of native plants. Ranches in western states often

Saskatchewan.91 In Wyoming, the federal managers of

preserve such mosaics, along with the open space that

the Thunder Basin National Grassland are working with

is valued by many groups. But the challenges of keep-

adjacent landowners and others to develop a plan for

ing such large areas intact are becoming greater. The

profitable livestock grazing even while the prairie dog

primary threats are plowing, the widespread effects of

population is restored to a level that is sufficient for the

oil and gas development, new wind farms and trans

reintroduction of black-footed ferrets. More collabora-

mission lines, and the desire of many newcomers to

tions of this nature are needed.92

build their homes in the middle of that open space. The

Working with interested landowners, various land

services provided by grasslands are reduced still further

trusts are finding ways to conserve the native grasslands

with a diminished landscape and the introductions of

that remain in a way that provides economic benefits to

exotic plants and animals that invariably occursome

the landowners and maintains biological diversity. Such

of which could become as problematic as sylvatic plague

projects may not have all the components of the original

or cheatgrass (see chapter 7). Various conservation

ecosystem, but they provide a means of conserving some

organizations (such as The Nature Conservancy and

of them. This approach is highly valued by conservation

other nongovernmental groups) are helping to main-

biologists. Indeed, it is the most feasible one over much

tain intact grasslands where landowners and agencies

of the West. Traditional approaches to livestock manage-

have common incentives for doing soeconomic or

ment are being adapted to achieve the goals of sustain-

otherwise.

ability and conservation. Notably, the circumstances

Another challenge is presented by climate change.

that led to livestock production on large ranches in the

As described in chapter 3, this has been occurring for

West have also, inadvertently, conserved open spaces

millennia, but the socioeconomic impacts now could

and much of the biodiversitythough not all of it.93

be substantial. Crop production could decline because

107

108 Plains and Intermountain Basins

of increased plant water stress and possibly less water

the season, such changes caused differences in nutrient

for irrigation. Forage could be reduced for large mam-

cycling and forage quality on the plots.97 For example,

mals of any kind. Further changes could occur if warm-

fringed sagewort, a small native shrub that is avoided

season grasses or shrubs become more common, which

by livestock, had much greater biomass than in the

have lower nutritive value than do cool-season grasses.

control plots; and there is evidence that less nutritious

For such reasons, the entire food web could change,

warm-season grasses will also become more common.98

belowground and above.

Research is under way to determine whether invasive

95

Notably, however, research in mixed-grass prairie

plants will be favored by elevated carbon dioxide as well

near Cheyenne indicates that increased atmospheric

as the nitrogen added by present-day levels of atmo-

carbon dioxide, while sometimes increasing the rate

spheric deposition.

of photosynthesis, has the additional effect of reduc-

Warming may also increase the rate of decomposi-

ing transpiration in some prairie plants. More water is

tion of soil organic matter, at least during times when

left in the soil for a longer time. A team of Colorado

the soil is moist, such as in the spring and on warm win-

and Wyoming scientists led by Agricultural Research

ter days. The overall effect could be a gradual decline

Service ecologist Jack Morgan conducted this research,

in the carbon stored in the soil with the concomitant

known as the Prairie Heating and Carbon Enhance-

release of more carbon dioxide to the atmosphere,

ment Experiment (PHACE). In replicated plots they

potentially aggravating global warming at a time when

increased levels of atmospheric carbon dioxide while

carbon sequestration is an important management

simultaneously warming the plant canopy and soil.

objective. Moreover, soils also could become more erod-

With higher amounts of carbon dioxide, the soil did

ible if plant cover declines; subtle changes could reduce

not dry as rapidly as in control plots, and plant water-

the capacity of grasslands to sustain grazing by live-

use efficiency was enhanced. This result is counter

stock. Animal scientists are already considering how the

intuitive, because if carbon emissions cause warming,

livestock industry might adapt its practices.

there should be higher rates of evapotranspiration,


causing more frequent periods of plant water stress.

Thus, the results of carbon dioxide enrichment and

However, as Morgan and his associates concluded, in

climate warming are more complicated than poten-

a warmer, CO2 -enriched world, both soil water content

tially causing higher rates of photosynthesis. If water

and productivity in semi-arid grasslands may be higher

stress develops quickly, or continues during extended

than previously expected. They added that, elevated

droughts, there will be no enhanced photosynthesis of

CO2 can completely reverse the desiccating effects of

the mixed-grass prairie, at least with the native plants

moderate warming.

that dominate the area now. The climate of the future

96

The PHACE study documented other changes as

may favor novel combinations of more drought-tolerant

well, including a rise in the importance of fungi, com-

species. Could this possibility be in the future for all

pared to bacteria, and an increase in the carbon-to-

kinds of ecosystems in the region? Will the ecosystems

nitrogen ratio of the soil. Along with more water later in

provide the services that many would like to have?

Chapter 7

Sagebrush

The iconic big sagebrush, Artemisia tridentata, dominates

Mention of sagebrush brings to mind wide-open

the most widespread ecosystem in Wyoming. Growing

spaces with few trees, few people, and scattered herds of

mostly west of the eastern grasslands, this shrub occurs

pronghorn and feral horses (fig. 7.1).2 Some would like

throughout the intermountain basins and covers about

to make the land more productive, whether by burn-

one-third of the state. Thats obvious to travelers. Not

ing to remove the shrubs or by cultivation. Indeed, the

so clear is that big sagebrush is but one of nearly a dozen

presence of knee-high big sagebrush is thought to be an

woody species of Artemisia that are commonly found

indicator of relatively deep soils that are suitable for crop

here. Others include alkali sagebrush, black sagebrush,

production. The potential for farming, however, is lim-

low sagebrush, silver sagebrush, and threetip sagebrush.

ited by scarce water and a cool, short growing season.

Fig. 7.1. Wyoming big sagebrush


steppe in central Wyoming,
southeast of Riverton. Associated
plants are junegrass, Sandberg
bluegrass, western wheatgrass,
blue grama, fringed sagewort,
phlox, and numerous other species. Scattered limber pine and
small groves of Douglas-fir occur
in the vicinity of Beaver Rim.
Elevation 7,200 feet.

109

110 Plains and Intermountain Basins


Fig. 7.2. Shrublands dominated by
Wyoming big sagebrush form a mosaic
with saltbush desert shrubland in this
area west of Rawlins. Big sagebrush is
restricted to ravines, where drifting
snow accumulates and provides more
water for plant growth. The adjacent
shrubland is dominated by birdfoot
sagewort, Gardners saltbush, western
wheatgrass, and winterfat. This area is
one of the driest in the state (see fig.
3.4). Elevation 6,750 feet.

The sagebrush-dominated ecosystem is far from uni-

ricegrass, pricklypear cactus, scarlet globemallow,

form. In some areas patches of sagebrush intermingle

horsebrush, and rabbitbrush. The common animals are

with patches of grassland, or with patches of small

often the same as well, including pronghorn, jackrabbit,

sagebrush plants intermingled with taller shrubs. This

badger, coyote, Wyoming ground squirrel, grasshopper

mosaic can sometimes be attributed to fires, which burn

and deer mice, several kinds of grasshoppers and spar-

unevenly across the landscape, creating patches of dif-

rows, and many others (see table 6.1).

ferent ages. The mosaic also is influenced by differences

Climate is an important factor in determining sage-

in soils and topography. Shallow soils and windswept

brush distribution and growth, as winter precipitation

ridges commonly have black sagebrush and small cush-

especially snowfallcontributes a larger proportion of

ion plants, whereas deeper, saline soils have birdsfoot

annual precipitation in the intermountain basins where

sage, Gardners saltbush, and greasewood. Soils along

big sagebrush is common (see chapter 3). 3 In contrast,

ephemeral creeks often support silver sagebrush. The

summer rains provide most of the water on the grass-

transitions are sometimes abrupt, due to abrupt changes

lands of the eastern plains. Unlike rainfall in the sum-

in moisture availability and soil conditions.

mer, which evaporates quickly, snow is more likely to

Many of the plants associated with sagebrush are

accumulate and when it melts, the water infiltrates

also found in grasslands, such as western wheatgrass,

to greater depths. This provides the more dependable

Sandberg bluegrass, needle-and-thread grass, Indian

water supply required by big sagebrush. In the Lara-

Sagebrush
Fig. 7.3. Mixed-grass prairie intermingles with Wyoming big sagebrush
steppe in the Powder River Basin. In this
area, the sagebrush cannot grow in the
ravines, probably because this habitat
is too wet for too long during the year.
Some of the shrubs on the edge of the
ravine are plains silver sagebrush, which
is more tolerant of wet soils. Elevation
4,800 feet.

mie and Shirley basins where there is less snowfall, or

comparatively moist ravines and valleys. It grows to a

toward the eastern fringes of its range, big sagebrush is

height of 6 feet or more, much taller than the other two

found almost entirely where snow drifting occurs, such

varieties. In Wyoming it is most often found in the Big-

as on the lee sides of ridges or in ravines (fig. 7.2). Nota-

horn Basin and the southwestern quarter of the state.

bly, big sagebrush is absent from very wet areas (fig. 7.3).

Veritable woodlands of head-high basin big sagebrush

Big sagebrush can be observed from the western

occur on stabilized dunes and other relatively moist

Dakotas all the way to the Cascade Mountains of Oregon

environments.5 Wyoming big sagebrush, much shorter

and Washington, and from Canada south to Arizona

than basin big sagebrush, is the most common shrub in

and New Mexico. It grows in habitats ranging from Utah

Wyoming. Normally less than 1.5 feet tall, it occupies

deserts at about 3,000 feet elevation to mountain shrub-

the drier uplands. The third subspecies, mountain big

lands at about 9,900 feet or higher. Surviving over such

sagebrush, is most common near woodlands of aspen

a broad range of environmental conditions requires

and low-elevation conifers but sometimes occurs in

considerable genetic variation. Three varieties with dif-

foothill ravines adjacent to shrublands with basin and

ferent adaptations to water stress are commonly recog-

Wyoming big sagebrush.6 Two of the subspecies can be

nized: basin big sagebrush, Wyoming big sagebrush, and

found growing together where environmental condi-

mountain big sagebrush.4 Basin big sagebrush is found

tions are intermediate. All three are known to hybridize,

at lower elevations (fig. 7.4) and is usually restricted to

which can lead to identification difficulties in the field.

111

112 Plains and Intermountain Basins

ELEVATION (feet)

10000

Mtn silver
sagebrush

Alpine sagewort

Mtn big
sagebrush

9000

Rawlins, black sagebrush is found on windswept ridges


or slopes with shallow soils and very little snow. On
lower slopes with more snow accumulation and deeper

Threetip
sagebrush

soils (but minimal salt accumulation), Wyoming big


sagebrush occurs. Mountain big sagebrush grows in
moist ravines with deep snow. Desert shrublands (see
next chapter) occur in dry basins, where snow depth is

Wyoming big
sagebrush

8000

accumulation is exceptionally deep, only grasses and

Black
sagebrush
or
Bud
sagewort
or
Birdfoot
sagewort

7000

Basin
big
6000 sagebrush
or
Silver
sagebrush
5000
Ravine
North
or
or
floodeast
plain
slope

moderate and salt accumulation is high.9 Where snow

South
or
west
slope

Ridge
or
dry
basin

MOISTURE GRADIENT
Fig. 7.4. Approximate distribution of different species of
Artemisia (sagebrush and sagewort) in relation to elevation
and topographic position.

forbs can survive (fig. 7.8). Clearly, sagebrush distribution in the landscape depends on soil moisture, salinity, depth, and texture, all of which vary over short
distances because of wind and topography.
Wyoming big sagebrush sometimes occurs in ovalshaped patches 1050 feet across. Commonly referred
to as sagebrush islands (fig. 7.9), some are simply patches
of tall sagebrush surrounded by shorter sagebrush.
Most are patches of sagebrush in a matrix of grassland
or desert shrubland. One type of island forms where
small silt dunes or sand dunes develop on the lee side
of taller shrubs. These dunes could have been initiated
by a period of wind erosion, possibly after heavy grazing by bison or livestock during a dry period. Where
an occasional shrub managed to survive or a new one
became established, sand and silt accumulated, creat-

Silver sagebrush, black sagebrush, and low sagebrush

ing conditions favorable for the growth and establish-

are also common. Silver sagebrush has two subspecies,

ment of other shrubs. With time, a larger number of

plains silver sagebrushfound on relatively moist sites

taller shrubs increased the deposition of windblown

at low elevations, especially east of the Continental

particles, including snow, resulting in the develop-

Divideand mountain silver sagebrush, which is com-

ment of small dunes that are 6 feet deep or more. Such

mon in moist meadows in the mountains throughout

dunes are known as coppice dunes and typically have

the region (figs. 7.5 and 7.6). Unlike most woody species

spiny hopsage, rubber rabbitbrush, green rabbitbrush,

of Artemisia, silver sagebrush is capable of sprouting from

and sometimes basin wildrye in addition to Wyoming

the root crown following fire or other disturbances.7

big sagebrush. Spiny hopsage is an indicator of sandy,

Black sagebrush is usually less than 18 inches tall and is

relatively moist soils in the semi-arid intermountain

found on relatively dry soils associated with the slopes

basins (due to the inverse texture effect, explained in

of escarpments or where a shallow hardpan of calcium

chapter 3). A desert pavement is sometimes found sur-

carbonate (caliche) develops. Similarly, low sagebrush,

rounding the islands, often with scattered plants of bud

which is usually less than 10 inches tall, forms patches

sagewort.10

on shallow soils interspersed with patches of big sage-

Elsewhere, the sagebrush islands are not associated

brush, such as in the lowlands of Jackson Hole and on

with windblown material and are less easily explained.

the Sweetwater Plateau. In the Bighorn Basin and other

In such places, the islands typically have numerous

areas of low precipitation, shallow soils commonly have

small mammal burrows, which can disrupt shallow

grasslands (fig. 7.7).8

hardpans or create soil conditions in other ways that

Another factor influencing the shrubland mosaic is

favor infiltration of water and taller shrub growth. Per-

snow drifting. For example, in the foothills south of

haps the animals are attracted by the shade provided by

Fig. 7.5. Approximate distribution maps for shrublands dominated by eight different
species or subspecies of sagebrush (Artemisia) in Wyoming. Adapted from Beetle and
Johnson (1982).

ELEVATION (feet)

5500
5000
4500
4000
3500

Fig. 7.6. Big sagebrush and silver sagebrush are easily distinguished by examining leaf shape. The longer leaves on big
sagebrush are ephemeral, dropping off early in the summer,
whereas the shorter leaves remain on the shrub for a full year.
Both species are evergreen. Drawing by Judy Knight.

10

20

30

SOIL DEPTH (inches)

40

Fig. 7.7.The distribution of grasslands (1), black sagebrush


steppe (l), and big sagebrush steppe ( ) in relation to soil
depth and elevation. Note how big sagebrush occurs on
deeper soils. Data from the Bighorn Basin.

Fig. 7.8. Wyoming big sagebrush in


the lowlands of northern Yellowstone
National Park. Nivation hollows form
where drifted snow persists on leeward
slopes until late June, creating a
meadow, because big sagebrush cannot
survive where the soils are wet during
much of the early summer. Elevation
7,400 feet.

Sagebrush

Other sagebrush islands are associated with mima


mounds (see chapter 9), or with pockets of deeper soil
in areas dominated by threetip sagebrush. As is usually the case, a vegetation pattern can have several
explanations.

Sagebrush Adaptations
Plant distribution is determined more by the tolerances of delicate seedlings than by the characteristics
of adult plants. Seedlings of big sagebrush are commonly observed in some years, almost always when
the early summer has been relatively cool and wet.12
Lower temperatures minimize the rate of water loss
from the drought-sensitive seedlings, providing more
time for roots to grow down to a dependable water
supply. Though the roots may not grow more than
an inch or two in the first year, the presence of deep
soil water contributes to a relatively moist surface soil
environ
ment in three ways: (1) capillary movement
from below, (2) condensation in the surface soil of
water vapor that emanates from the deeper soil water,
Fig. 7.9. Wyoming big sagebrush sometimes grows in patches,
such as in this area east of Rock Springs. The patches can be
small or large, oval or irregular, and often are associated with
coppice dunes. Elevation 7,000 feet.

and (3) hydraulic redistribution from deep to shallow


soil via the root system of other plants (see below for
explanation). Once established, the adult plants tolerate the occasional years when little precipitation falls
during the winter. Surface roots use surface moisture

the shrubs. But how did the shrubs become established

early in the growing season, but other roots grow

in the first place? One explanation is that the patch

down to 6feet or more and make use of deep soil mois-

began with a single shrub that survived by chance,

ture aftersurface soils are dry. The root system extends

perhaps adjacent to a burrow. Its presence caused snow

laterally to a distance of about 5 feet, well beyond the

drifting on the lee side of the shrub, creating condi-

shrubs canopy.

tions favorable for the seedlings of additional shrubs

Significant soil water recharge at depth usually

perhaps from seed produced by the original plant. In

occurs only once a year, during the snowmelt period,

addition to snow, the taller shrubs could also lead to

but, as noted, surface soil dampening during a hot, dry

the accumulation of blowing organic matter and soil

summer can happen even without rainfall through

particles. This accumulation, combined with nutrients

the novel process of hydraulic redistribution.13 First

brought to the surface by the animals and litterfall from

described by ecologists James Richards and Martyn

the shrubs, creates islands of fertility. Such islands are

Caldwell in 1987, this process is now known to occur

found where water and other resources are inadequate

in numerous plants that have both deep and shallow

11

to support a uniform cover of sagebrush. Neil West, an

roots. If the deep roots are in moist soil and the shal-

ecologist at Utah State University, suggested in 1988

low ones in dry soil, as often happens in the summer

that, for sagebrush-dominated vegetation in the Great

where big sagebrush grows, then water is transported

Basin of Utah, any activity that destroys sagebrush

by the deep roots up to the shallow roots, where it is

islands reduces the productivity of the ecosystem as a

released into the soil. Initially the process was referred

whole. The same probably is true for Wyoming.

to as hydraulic lift (from deep to shallow soil), but sub-

115

116 Plains and Intermountain Basins

sequent research found that water can be moved up or

The carbohydrates produced by photosynthesis are

down (always from wet to dry soil).14 This process pro-

distributed to all plant parts where energy is needed. In

vides more water in the summer for plants that have

big sagebrush, a large proportion of these energy-rich

only shallow roots, thereby extending the growing

compounds is stored in the twigs. Thus, both the twigs

season and increasing the overall productivity of the

and the leaves are a good source of food for herbivores.

sagebrush ecosystem. Hydraulic redistribution also can

Furthermore, the shrubs are tall enough to extend above

prolong microbial activity in the surface soil, which

the snow during most winters. Little wonder, then, that

facilitates decomposition and increases the supply of

historically the shrubs have been subjected to consid-

nutrients for the plants.

erable browsing pressure and that they have evolved

Several mechanisms allow the efficient use of water

a mechanism to minimize this herbivory, namely, the

by big sagebrush. The stomata close rapidly as water

terpenes that give sagebrush its characteristic aroma.

stress develops during the day, limiting transpiration

The terpenes apparently cause sufficient indigestion to

until the next morning, when they are again open.15

minimize the amount eaten by some animals. Without

The plant re-equilibrates at night, when water uptake

this adaptation, the shrub may not have survived the

exceeds losses. Stomatal closure limits photosynthesis

intense browsing that might have resulted from large

as well, but prolonging the growing season by con-

herbivore populations, especially those limited most

serving water seems to be adaptive. In addition, the

directly by winter food supply.18

plant is evergreen, retaining most of its leaves for one

Considering the genetic variation that exists in popu-

year. Moreover, acclimation to changing temperatures

lations of widespread plants, it should not be surprising

occurs readily, even down to freezing. Big sagebrush is

that different species of sagebrush, and even different

capable of photosynthesis in the early spring and even

individuals of the same species, vary in the amount of

during relatively warm winter days.16 In general, ever-

terpenes they produce. Herbivores have been observed

green plants fix carbon whenever water and tempera-

to preferentially browse on those plants with the low-

ture conditions are within tolerance ranges. With no

est terpene concentrations. Some animals tolerate the

time lost for the production of new leaves, the growing

terpenes. A good example is the pronghorn, which

season is longer.

eats sagebrush all year long (the shrub provides 7095

Water is most readily available in the spring, when

percent of its food in the winter). Other herbivores eat

big sagebrush has the largest leaf area for photosyn-

sagebrush as well, such as sage-grouseespecially when

thesis. Two leaf types are produced: ephemeral leaves

other foods are not available.19

in early spring, which fall as water stress develops

Big sagebrush, unlike most of its neighboring plants,

during the summer, and overwintering leaves that

lacks the capacity to sprout from roots or the root crown.

remain on the plant until the following spring (see fig.

Consequently, longevity and seed production are espe-

7.6).17 Essentially, the plant is evergreen, but the loss

cially important for its persistence. Some plants live for

of ephemeral leaves by early summer allows for some

100 years or more, though 4050 years is more com-

reduction in total leaf area as water becomes more

mon. 20 The shrub produces annual rings in the wood

limiting, providing another mechanism for conserv-

similar to temperate zone trees. With regard to sexual

ing water and extending the length of the growing

reproduction, big sagebrush produces thousands of tiny

season. Shadscale, another shrub in the basins, also

seeds each year, some of which remain viable for up to

produces both ephemeral and persistent leaves. The

4 years.21 Like many other plants, germination occurs

leaves of many species of sagebrush and other xero-

in late winter or early spring, when moisture is avail-

phytic plants often have a pale, silvery-green color,

able and the temperature is warm enough for growth.

attributable to dense, minute silvery hairs on the sur-

No cold period is required to break seed dormancy,

face. This color reflects light more readily than dark

although some investigators have found that this treat-

green does, which reduces water loss because the

ment causes higher germination rates in mountain big

leaves remain cooler.

sagebrush. Even very low salt concentrations can inhibit

Sagebrush

big sagebrush germination,22 an observation that may


explain why the plant is less common on saline soils.
A high percentage of big sagebrush seeds are viable
and germinate readily (80 percent or more). However,
seedlings survive only during favorable years and typically following a disturbance that reduces the competitive ability of neighboring plants.23 The age distribution
of a sagebrush population is therefore discontinuous; that
is, only a few age classes are represented.24 Conditions for
high growth rates early in the summer are essential, so
that the seedlings can develop a root system adequate to
cope with late-summer droughts. Jerry Schuman, mineland reclamation specialist, found that sowing too much
grass seed can lead to a level of competition from grasses
that reduces sagebrush seedling survival.25

The Sagebrush Ecosystem


Compared to grasslands, the distinguishing features
of sagebrush ecosystems are (1) the presence of an aromatic shrub with both shallow and deep roots and (2)a
large portion of the annual precipitation occurring in
the winter. Otherwise, grasslands and sagebrush steppe
are similar:
plant growth (net primary productivity) is limited

Fig. 7.10. Snow frequently accumulates on the lee side of


shrubs, which probably increases the amount of water that
infiltrates the soil.

by water availability and the length of the growing


season;
most of the biomass and half or more of the herbivory is belowground;
potential evapotranspiration usually is greater than
the annual precipitation;
nutrients are rarely leached beyond the rooting zone;
fire, drought, and burrowing animals are common
disturbances;
plant and animal species have evolved to minimize
competition by using different resources and by us
ing them at different times of the year; and
coexisting herbivores, above- and belowground,
interact in complex ways to affect nutrient cycling
and plant growth.

Hydrology and Plant Growth


Drifted snow on the lee side of shrubs (fig. 7.10) can be a
significant supplement to soil water recharge, especially
on comparatively level sites, where accumulation is not
affected by topography.26 Scientists and land managers
have learned that, without big sagebrush, relatively
more water and nutrients are available for the growth
of more palatable grasses and forbs, but overall, there
is probably less total water available and less total plant
growth than when the shrubs are present. Sagebrush
ecosystems have aboveground plant growth totaling
80250 g/m2/year, which is higher than that for many
mixed-grass prairies in Wyoming.27
Plant water availability may also be influenced by

This section focuses on differences caused by the pres-

the blackbody effect of shrubs that are above the snow

ence of an evergreen, deep-rooted shrub that is not able

in winter.28 Heat is absorbed by the shrub tops, caus-

to sproutand what happens when it dies because of

ing snowmelt around the shrub more often than would

burning or herbicides.

occur otherwise and creating a depression or well into

117

118 Plains and Intermountain Basins

which more snow drifts during the next storm. In this

if a fire occurs. Big sagebrushdominated ecosystems

way, the shrubs facilitate the percolation of meltwater

burn from time to time, and because the sagebrush

deeper into the snowpack or possibly into the soil (if its

itself cannot sprout, nutrient and water resources are

not frozen), reducing the amount that is lost to subli-

made available to grasses and forbs for 10 years or

mation while creating the potential for additional snow

more. As herbaceous plants become more abundant,

accumulation as the snow melts around the shrub tops.

the carrying capacity for bison and cattle increases. Per-

The depressions around the shrubs can be created and

haps the optimal habitat for some animals is a mosaic

filled several times during a winter, augmenting mois-

of old-growth sagebrush intermingled with recently

ture input above that to be expected if the shrubs were

burned sagebrush. Before the arrival of EuroAmericans,

not there. This phenomenon could be especially impor-

the steppes dominated by sagebrush might have been

tant where the vegetation is more or less uniformly

more varied where periodic fires were possible, creat-

covered with snow except for the tops of shrubsa

ing patches of grassland and young sagebrush inter-

common occurrence.

mingled with old-growth sagebrush. Today, sagebrush

As noted, big sagebrush has roots that are deeper

cover might be more uniform and older in some places

than those of most grasses and forbs. As a result, more

than previously, the result of fire suppression and more

water is used by plants when the shrub is present. If the

extensive livestock grazing.32 However, elsewhere, sage-

shrubs are removed by fire or other means, total water

brush landscapes have been heavily fragmented, as dis-

consumption by the plants (transpiration) is less, per-

cussed later. Because of threatened species that depend

haps by 15 percent,

on sagebrush, the practice of prescribed burning in this

29

30

but total plant growth probably is

lower as wellas would be expected based on the cor-

ecosystem may not be advisable in some areas. 33

relation of plant growth with evapotranspiration in arid

Total annual plant growth is enhanced by the pres-

climates (see fig. 6.5). Closer to the mountains, where

ence of big sagebrush, in part because of its deeper root

snowfall is higher, less transpiration may allow for some

system but also because its green leaves persist through-

runoff as streamflow.

out the year, which lengthens the growing season. 34


Sagebrush growth tends to be greatest in the fall and

Energy, Carbon, and Soil Organic Matter

early spring, whereas the growth of associated grasses


and forbs is more rapid in late spring and early sum-

The food web of sagebrush-dominated shrublands is

mer. Fall-through-spring precipitation and temperature

similar to that of grasslands, with probably more than

are good predictors of total annual growth of sagebrush

half the herbivory and carnivory occurring in the soil.

and associated plants.35

However, less than half the annual precipitation comes

The flow of carbon through ecosystems parallels

during the summer, which limits the success of warm-

the flow of energy, because so much energy is stored

season C4 plants, such as blue grama. Much of the plant

as carbon-rich compoundsthat is, organic materials,

growth is concentrated in the less-palatable big sage-

such as cellulose and other carbohydrates. They are

brush, which competes for moisture that could be used

important sources of food for large and small herbivores,

by other forage plants, namely, cool-season grasses.

aboveground and below. Moreover, animal carcasses

Consequently, the forage available for large ungulates

along with plant remains that do not burn become part

might be less than on the grasslands, especially those

of the mulch and soil organic matterboth of which

on the Great Plains where the plants have evolved with

increase water infiltration and nutrient availability. Big

more bison herbivory (see chapter 6).31 Sagebrush is

sagebrush tends to concentrate such materials around

known to be an important food for pronghorn, mule

its base more than do herbaceous plants, creating a

deer, and elk, but primarily during the winter, when

small-scale island of fertility that can persist even after

herbaceous forage is scarce or buried by snow. Bison rely

sagebrush is burned or removed in some other way.36

on grasses and forbs more than shrubs, as do cattle.

This patchiness in the surface soil helps promote sage-

Just as in other ecosystems, the energy stored in

brush ecosystem recovery after disturbances. In general,

stems and leaves aboveground can be released as heat

mineland reclamation specialists have found that con-

Sagebrush

serving as much soil organic matter as possible, along

other plants, such as Indian ricegrass.39 Losses of nitro-

with its associated microorganisms, is fundamental for

gen may occur through erosion, denitrification, or the

restoration.

emission of nitrogen gases.40

Because of concerns about the effects of carbon di

Several investigators have studied the rates of litter

oxide on climate change, ecologists have been conduct-

fall and decomposition in shrublands dominated by

ing experiments to determine whether the sagebrush

big sagebrush, both important processes for nutrient

ecosystem is a source of atmospheric carbon or a sink:

cycling.41 Some evidence suggests that nutrients are

does an ecosystem with big sagebrush tend to store

reabsorbed by twigs before the leaves fall, thereby con-

more or less carbon than is released from the ecosystem

serving nutrients in the shrubs.42 Still, soil nitrogen

as carbon dioxide?37 As with most semi-arid shrublands

is enhanced under shrubs. There also is great varia-

around the world, sagebrush steppe tends to be a sink

tion in soil organic matter and nitrogen availability

during wet years but a source in dry years. Generally,

higher in ravines with mountain big sagebrush to less

the sagebrush ecosystem accumulates carbon as the

on ridgetops with black sagebrush.43 Burrowing mam-

shrubs age, in the form of organic matter above- and

mals and harvester ants affect nutrient dynamics as

belowground. When most of the shrubs die, the eco-

well.44

system could become a source if the unburned woody

The major effect on nutrient cycling of having big

sagebrush biomass decays rapidly. However, Wyoming

sagebrush or any shrub in the ecosystem is to have more

ecologists Meagan Cleary, Elise Pendall, and Brent

woody biomass aboveground. Nutrient accumulation is

Ewers did not detect an increase in carbon emissions

therefore greater aboveground than where grasses and

after firepossibly because the growth of grasses and

forbs predominate. When the shrub dies, or as twigs

forbs increases greatly soon after fires, and the sage-

and branches die, more of the nutrients are added to

brush roots and stems decompose slowly.38

the soil surface as wood rather than as herbaceous plant

Still, the benefits of improving habitat or forage availability for some species by burning must be balanced

material. Also, woody fuels accumulate that can create


conditions for a longer, hotter fire.

against the undesirable effects of additional atmo-

Another influence of shrubs is to create islands of

spheric carbon dioxide from the fire, not to mention

fertility, as discussed previously. Shrubs accumulate

the invasion of undesirable plants (such as cheatgrass)

windblown debris and associated nutrients, and the

and the loss of critical habitat for threatened species (see

root system of the shrubs extracts soil nutrients from

below). Federal and state guidelines often call for man-

several feet beyond the canopy while dropping them on

aging ecosystems on public lands so that they sequester

the surface as litter in a smaller areapossibly another

carbon, resist the invasion of exotic plants, and offer

mechanism creating patchiness in the vegetation. The

habitat for such species as the greater sage-grouse and

death of the sagebrush root system must be important

pygmy rabbitall ecosystem services that are now

as well, not only in reducing competition for water and

highly valued.

nutrients with neighboring grasses and forbs, but also


in providing a substantial pool of nutrients as the roots

Nutrient Availability

decompose. The decaying root system can also alter soil


structure, possibly leading to greater infiltration rates.

Nutrient distribution in sagebrush-dominated eco

Such effects may be lost at some point after big sage-

systems is similar to that of grasslands, with the largest

brush has been removed, whether by fire or herbicides.

amounts in the soil. Water is usually more limiting to

Deeper soil organic matter may gradually be lost as well,

plant growth than are nutrients, though nitrogen may

along with the potential for hydraulic lift and the tap-

be limiting during wet years. As in grasslands, the major

ping of deep soil nutrients, including phosphorus.45 Of

source of nitrogen is probably precipitation, but symbi-

course, big sagebrush typically becomes re-established,

otic nitrogen fixation occurs locally in biological crusts

as discussed in the next section, restoring the benefits

on the soil surface; in the nodules of lupine and other

of having this shrub in the ecosystemunless cheat-

legumes; and possibly adjacent to the roots of some

grass changes everything.

119

120 Plains and Intermountain Basins

Recovery after Disturbances


Big sagebrush is easily killed because it cannot sprout,
but historically it re-established after disturbances.
The rate of recovery varies considerably in relation to
the amount of precipitation, the amount of litter on
the soil, the degree of competition from herbaceous
plants, the intensity of grazing, the number of sagebrush seeds in the soil, the number of live shrubs that
remain, and the kind of big sagebrush on the site.46
Mountain big sagebrush has been observed to reestablish after about 40 years, whereas Wyoming big
sagebrush takes much longer.47 Without big sagebrush,
the ecosystem appears and functions like a grassland,
with junegrass, needle-and-thread grass, western
wheatgrass, Sandberg bluegrass, and numerous other
species. In relatively moist environmentsand in the
absence of cheatgrasssagebrush cover becomes quite
high as the shrubs age, persisting until the next fire or
drought.48
Even without planting, managers have come to
expect that big sagebrush and other native plants will
eventually invade areas where the plants have grown
before. That happened, for example, on abandoned
farmland in the Powder River Basin.49 Interestingly, the
establishment of big sagebrush during mineland reclamation is often difficult. Some research suggests that
survival is higher when mycorrhizal fungi are available
in the topsoil and when the recommended seeding rate

western Wyoming and neighboring states.53 In contrast,


shrub dieback during the mid-1980s in the Great Basin
was attributed to unusually wet conditions that persisted
for 5 years.54 Shrubs that were affected included antelope
bitterbrush, big sagebrush, fourwing saltbush, rabbitbrush, shadscale, and winterfat.
Extreme, ecologically significant weather events are
likely to become more frequent with climate change.
Higher mean annual temperatures will lead to less snow
accumulation and earlier snowmelt, and also a higher
potential for evapotranspiration throughout the year.
Because of added warmth, drier conditions for longer
periods during the summer could result even if annual
precipitation stays the same. If plants begin growing earlier in the spring, late-spring frosts could kill
some of them. Predicting precipitation changes with
global warming is currently difficult, but if droughts
are more severe, some shrublands now dominated by
sagebrush could become dominated by more droughttolerant shrubs or possibly grasslands. 55 Silver sagebrush
might be affected most severely, but big sagebrush also
depends on higher levels of precipitation than do most
grassland species. Sagebrush reinvasion after disturbances probably would be slowed, if it occurs at all,
because the extended cool, moist conditions required
for seedling establishment may become less frequent.
Drought-tolerant invasive species, such as cheatgrass,
may have a better chance of becoming established.

of perennial grasses is not too high.50

Grasshoppers
Drought, Spring Frost, and Extended Wet Periods
Unusual weather can kill shrubs, though the degree of
disturbance depends on the species. In a study in eastern
Montana, the shoots of silver sagebrush were more susceptible to drought than were those of big sagebrush, but
the silver sagebrush recovered more rapidly because of its
ability to sprout from the root crown.51 Winter mortality
of sagebrush can occur as well, probably from water stress
created by frozen soils, low soil water availability, and
below-average snowfall.52 Another cause may be the premature breaking of dormancy during periods of aboveaverage air temperature, making the plants susceptible to
subsequent frosts. Extensive areas of mountain big sagebrush have been killed by spring frost damage in south-

Outbreaks of grasshoppers killed approximately 50percent of the big sagebrush in a portion of the Powder
River Basin during the drought years of the 1930s.
Allred wrote, the grasshoppers swarmed in such hordes
that they devoured all of the edible vegetation, ate the
leaves and bark from the twigs of the sagebrush, and
completely girdled the more tender stems.56 Other
insects known to kill big sagebrush include gall midges,
the aroga moth, some beetles, and Mormon crickets.
Small mammals (such as voles) may girdle the stems of
big sagebrush, causing a significant disturbance when
their populations are high.57
Grasshopper outbreaks tend to occur during drought
years that are relatively warm. Thus, continued warm-

Sagebrush

ing of the climate could favor grasshoppers, killing

elevations, where mountain big sagebrush is the domi-

many of the shrubs and reducing forage availability

nant shrub.61 Fire rotation can be thought of as the aver-

at a time when it already is in short supply because of

age time required for one or more fires to burn an area

drought. Historically, grasshopper densities fluctuate

equal to the entire landscape under consideration. Its

greatly from one year to the next, leading entomolo-

tempting to think that global warming and drought will

gists to conclude that control measures in most parts of

shorten the fire-return interval in sagebrush ecosystems.

Wyoming are beneficial only during the year that insec-

However, if droughts are severe enough, the interval

ticides are applied. Usually, little or no multi-year bene-

could be lengthened because of less fuel production.

fit occurs.58 It is still unknown whether global warming

Gaining confidence in fire rotation estimates is im


portant, because prescribed fire can be a good manage

will change that conclusion.

ment strategy in some places. However, at the present


time, how often should sagebrush be burned, if at all?

Fire

Has fire suppression really created a stand in need of res-

Once started, whether by humans or lightning, fire can

toration, or might the area be thought of as old-growth

burn large areas of sagebrush. In 1868 James Chisholm

sagebrush with values for wildlife? Answers to such

explored the Wind River Basin and the vicinity of

questions are required for the proper management of

South Pass City during the Wyoming gold rush. He

animals that are sagebrush obligates. As the sagebrush

observed the rapid spread of at least two human-caused

habitat is fragmented in one way or another, these

fires across what must have been sagebrush-dominated

species have less habitat available to them, potentially

uplands,59 writing in his journal:

reducing their populations to the point where they

The grass took fire and all our efforts could not extinguish it. This time the situation was really alarming,
for there was a prospect that the entire Wind River
Valley might go up in a flame. . . . The flame went over
the nearest hill with amazing velocity, Heaven knows
how far. . . . The mountains were black and bare over
which we travelled for the rest of that day, and we saw
the fire pursuing its way far ahead in several directions, but fortunately away from the valleys.

become locally rare or extinct. Significantly, the reestablishment of sagebrush can require several decades,
and if invasive plants (such as cheatgrass) already have
seeds in the soil, there is the potential that sagebrush
may not recover at all.
In general, the effects of fire on sagebrush eco
systems are more prominent than those following
grassland fires because many of the shrubs are killed
(figs. 7.11 and 7.12). Other native species also may
be reduced in abundance by burning, such as Idaho

The words mountains and away from the valleys

fescue, needle-and-thread grass, pricklypear cactus,

suggest that much of the burned area was in the foot-

and threadleaf sedgethough others can be favored,

hills or higher, where mountain big sagebrush probably

such as bottlebrush squirreltail, Sandberg bluegrass,

was the dominant shrub and fuel accumulation might

and western wheatgrass.62 Although the cover of big

have been greater.

sagebrush is greatly reduced by fire, several sprout-

The mean fire-return interval in sagebrush steppe

ing shrubs persist or increase in some locations, for

in the 1800s is thought to have ranged from 20 to 100

ex
ample, horsebrush, rabbitbrush, and winterfat in

years, depending on the potential for plant growth and

relatively dry basins; and antelope bitterbrush, prairie

fuel accumulation on the site.

However, University of

rose, silver sagebrush, skunkbush sumac, and western

Wyoming ecologist William Baker did a thorough review

snowberry in more moist environments.63 Sprouting

of the methods used for calculating such estimates and

species definitely are favored if fires occur more than

concluded that fire rotations probably were much longer:

once every 2025 years. As expected, perennial grass

325450 years in shrublands dominated by low sage-

production commonly increases when competition

brush; 100240 years in shrublands dominated by Wyo-

from shrubs is reduced.64 Unfortunately, undesired

ming big sagebrush; and 70200 years or more at higher

weedy species also can increase.65

60

121

122 Plains and Intermountain Basins


Fig. 7.11.The growth of grasses and
forbs increases greatly after big sagebrush is reduced by burning, as in this
area northwest of Rock Springs. The
vegetation on the left provides more
forage for livestock but is less favorable
for sage-grouse. Burning sagebrush can
favor the establishment of cheatgrass
and other invasive plants (see fig. 7.13).

For herbaceous plants in the sagebrush ecosystem,

a critical resource, whether for livestock or big game.

susceptibility to fire-caused mortality seems to depend

Burning sagebrush may be useful, but it should be done

on the amount of fuel that has accumulated above the

only in the context of management plans and after

root crowns, the depth of the latent buds below the soil

careful consideration of the undesirable changes that

surface, and the stage of growth at the time of burn-

could result, including the spread of invasive plants and

ing. Some research suggests that western wheatgrass

undesirable effects on sagebrush obligate animals, such

vigor and production increase with spring burning but

as sage-grouse.67

decline with fall burning, and that burning at any time


reduces the productivity of needle-and-thread grass.66
Repeated burning in consecutive years can cause a

Cheatgrass

shift from a community dominated by both cool- and

Though problematic for some sagebrush-dependent

warm-season species to one dominated primarily by

animals, prescribed fire simulates a natural distur-

warm-season species, which are generally more fire

bance to which many native plant species are adapted.

tolerant. This shift in dominance can be significant to

Older shrubs are removed and herbaceous production

land managers if early-season production is viewed as

increases. Unfortunately, invasive plants also may ben-

Sagebrush

300

Grasses

PERCENTAGE

200

a winter annual that begins growth in the fall or early

Thickspike
wheatgrass and
Plains reedgrass

Bluebunch
wheatgrass

spring and gains a competitive edge over many native


species that are slower to initiate growth. Generally,
fire and heavy livestock grazing hasten the spread of
cheatgrass, though some research suggests that proper

Bluegrass

100

management of native perennial species can slow this


invasion.72 Native species are more likely to become
Idaho fescue

300

Needle-and-thread

established if actions are taken to favor such species


within a year or two of a fire, for example, by seeding

Shrubs

Horsebrush

desired native species and protecting the rangeland

Rabbitbrush

200

from excessive grazing. The maintenance of biological


soil crusts seems to be important as well (as discussed
later).

100

With cheatgrass as a component of the flora (or even


just in the seed bank), land managers are confronted

Sagebrush

1936

1942

with the conundrum of increasing desired native

1948

1954

YEAR

1960

1966

Fig. 7.12.Changes in the canopy cover of several grasses and


shrubs after the burning of big sagebrush steppe in eastern
Idaho. The burn occurred in 1937. An increase (or decrease)
in plant cover is indicated when the line for a species is above
(or below) the dashed line. The two shrubs that increased, rabbitbrush and horsebrush, are capable of sprouting from their
roots; big sagebrush declined, because it does not sprout, but
it eventually became re-established from seed. Adapted from
Harniss and Murray (1973).

grasses by burning big sagebrush without increasing the


abundance of an undesired, introduced weed.73 Moreover, cheatgrass can become so abundant that it greatly
increases the flammability of the vegetation. Fires then
occur more frequently, diminishing the chances of sagebrush re-establishment and causing a decline in some
perennial grass speciesfavoring cheatgrass expansion
still further (fig. 7.13).74 The economic impact of cheatgrass is moderated to some degree by the fact that it can
be good forage before it flowers, but converting semiarid steppes to vegetation dominated by an introduced

efit, some becoming so abundant that they reduce the

annual is a significant change that has undesirable

growth of native species and the potential for restor-

consequences. In addition to increasing flammability,

ing wildlife habitat. Cheatgrass (Bromus tectorum), also

cheatgrass is associated with an increase in soil temper-

known as downy brome, is the most notorious example.

ature and erosion, along with reductions in soil organic

This annual plant has greatly complicated the practice

matter and small mammal abundance and diversity.75

of prescribed burning as a management tool. For many

The cheatgrass problem is not restricted to land man-

years managers thought that cheatgrass would not grow

aged for livestock. For example, a fire burned a large stand

well at higher elevations, but it is now commonly found

of big sagebrush in Little Bighorn Battlefield National

throughout Wyoming, even at elevations above 7,000

Monument in southern Montana, where livestock have

feet.68 With climate warming, it very likely will become

not grazed for many years. Much of the sagebrush was

more common in some parts of the state.69 Other prob-

killed, and bluebunch wheatgrass, a co-dominant before

lematic species in shrublands include broad-leaved

the burn, became less abundant. Cheatgrass became

pepper
weed, Canada thistle, Dyers woad, halogeton,

more common on the burned area, as did yellow sal-

leafy spurge, Russian knapweed, spotted knapweed,

sify. Managing the vegetation of a national monu-

Russian thistle, western salsify, and yellow and Dalma-

ment so that it reflects conditions prior to the arrival of

tian toadflax.70

EuroAmericans is a goal that sometimes seems impos-

Cheatgrass was introduced to western North Amer-

sible once invasive species become established. Further

ica in the late 1800s from the steppes of Europe.71 It has

research is necessary to resolve perplexing ecological

become common in many areas, because the species is

problems such as this one.

123

124 Plains and Intermountain Basins

Fire suppression
Fire

Native perennial grassland

Sagebrush steppe with perennial grasses

Heavy grazing,
particularly
during spring

Moderate
grazing

Possible recovery
if no fire and only
fall sheep grazing

Fire
Native grass cover with possible
shift toward less-palatable warmseason species; rabbitbrush and
sagebrush may increase slightly

Only annuals,
rabbitbrush, and
sagebrush
Fire again
Only annuals;
increased erosion

Possible recovery
with reduced grazing
if no significant loss
of soil and no fire

More
frequent fire

Near-permanent
deterioration of site

Cheatgrass and other annuals increase, as do sagebrush,


rabbitbrush, and less-palatable grasses and forbs

Fig. 7.13.The apparent effect of cheatgrass introduction,


heavy livestock grazing, and increased fire frequency in the
big sagebrush steppes of Utah and Nevada. A similar scenario

may apply to Wyoming if cheatgrass becomes more common.


The introduction of exotic species can cause undesirable ecosystem changes. Adapted from West (1988).

A confounding factor may be climate change, as

be the best approach, along with adjustments in live-

both warming and higher levels of carbon dioxide

stock management, but that may require highly dis-

would probably favor cheatgrass expansion. Even

ruptive weed control practices, and then resting the

without warming, annual plants often evolve rapidly

rangeland for many years while native species become

to tolerate new environmental conditions. In addition,

re-established.76 Restoring biological soil crusts may be

other disturbances caused by road building, plowing,

required, about which little is known.

and various industries have created new habitats for

One way to stop or slow the spread of cheatgrass

this invader and others. Of great concern, cheatgrass

is to suppress fires where cheatgrass invasion is likely

an annualpersists even with very little moisture.

to occur, at least until effective and affordable control

After a decade or more of cheatgrass occupancy and

strategies are found. Curiously, there is evidence for the

the concomitant reduction of native species, the land

intriguing idea that some native species may be evolv-

could be left with little plant cover. Soil erosion would

ing adaptations for competing with invasive species.

accelerate.

However, preventing the invasion of exotic plants and

Efforts to control cheatgrass have generally been

pathogens should be a high priority where that is still

futile. Experiments with herbicides, the introduction

an option and where there is a desire to maintain bio

of pathogens, and strategically timed fires and grazing

diversity and the services provided by native ecosystems

have had some success, but only at considerable cost.

(see chapter 6 for further discussion of invasive plant

Promoting the establishment of native species may

management).77

Sagebrush

Effects of Livestock and Feral Horses on


Sagebrush Ecosystems
As discussed for grasslands, grazing is not a disturbance
unless too many animals are confined in too small an
area for too long a time. Bison, pronghorn, elk, deer,

this plant. This climate has been found very favorable to the restoration of health, particularly in cases
of consumption; and possibly the respiration of air
so highly impregnated by aromatic plants may have
some influence.79

and other herbivores were integral parts of the sage-

Fremonts cartographer, Charles Preuss, wrote, When

brush ecosystem for thousands of years, and they still

the wind blows, it is as though one were in a pharmacy.

are in some places. Ungulate grazing does not generally

Ten years later, in 1852, Platt and Slater wrote, Along

kill the dominant big sagebrush, though excess herbiv-

the Sweet Water, most of the way, are narrow bottoms

ory is known to cause declines in plant cover and, in

of good grass. Adjacent to these bottoms are large, arid,

some places, an increase in undesired plants.

wild-sage plains, extending to the mountains.80

One of the most widely mentioned responses to live-

The same is true today (fig. 7.14). Big sagebrush clearly

stock grazing is an increase in big sagebrush cover. How-

has been a dominant feature of the Wyoming landscape

ever, sagebrush abundance should not be considered an

for hundreds of years. Its abundance might have been

artifact of livestock grazing pressure everywhere.78 Long

favored by grazing in some areas, because the removal

before the cattle drives from Texas, John Fremont wrote

of palatable species reduced competition for sagebrush

about his travels across Wyoming in 1842:

seedlings. Such trends might have been most common

One of the prominent characteristics in the face of


the country is the extraordinary abundance of the
artemisias. They grow everywhereon the hills,
and over the river bottoms, in tough, twisted, wiry
clumps; and, wherever the beaten track was left, they
rendered the progress of the carts rough and slow. As
the country increased in elevation on our advance to
the west, they increased in size; and the whole air is
strongly impregnated and saturated with the odor of
camphor and spirits of turpentine which belongs to

on the fringes of the Great Plainsfor example, in portions of the Powder River Basin, where summer rainfall
is favorable for grasses and forbs.81
Livestock grazing can have dramatic effects on the
grasses and forbs associated with big sagebrush, but
drought is another important factor. Several studies
have found that conditions appearing to have resulted
from poor livestock management were actually caused
by extended dry periods.82 Indeed, short-term shifts in
climate can cause great changes in plant growth and

Fig. 7.14.This photograph


along the Sweetwater River
east of Jeffrey City was taken
by William Henry Jackson
in 1870. The vegetation,
dominated by Wyoming big
sagebrush, is essentially the
same today (Johnson 1987).
The granitic Sweetwater
Rocks on the left have limber
pine and Rocky Mountain
juniper. Elevation 6,200 feet.
U.S. Geological Survey photo
JWH00292.

125

126 Plains and Intermountain Basins

the best examples of soil crusts in Wyoming have been


found in the relatively warm, desert-like parts of the
Bighorn Basin (fig. 7.15).
Because of the dramatic changes in the sagebrush
ecosystem that can be brought about by poorly managed livestock grazing, some ecologists have concluded
that, in contrast to grasslands on the Great Plains to the
east, the plants of the sagebrush steppe did not evolve
with grazing by bison and other large mammals.85 The
rationale is that the environment was unfavorable for
large herds of bison because of summer drought and the
scarcity of warm-season C4 species at that time.86 Therefore, food to sustain lactating cows was less available
during the summer. Biotic crusts might have formed
under these conditions, where bison populations were
low. When cattle and sheep became abundant in the
late 1800s, trampling would have destroyed the crusts,
causing the soil to become more erodible and more suitable for the invasion of introduced weedy annuals, such
as cheatgrass.
But does this scenario apply to the higher elevation
landscapes of Wyoming, where the accounts of early
explorers suggest that bison were common in sageFig. 7.15. Microphytic soil crusts cover the soil between shrubs
in this grazing exclosure in the Bighorn Basin east of Greybull. The common shrub in the exclosure is winterfat, which
is characteristic of desert shrublands (see chapter 8). Soil crusts
tend to occur more often in drier areas and are less common
in Wyoming than in the Great Basin to the west. Photo by
Jennifer Muscha.

brush-dominated landscapes? During midsummer in


1832 and 1833, Nathaniel Wyeth observed buffalo by
the hundred thousands in the vicinity of South Pass
and Big Sandy Creekwhere sagebrush would have
been abundant. At one point he noted that the country was covered with buffalo. In the same area, Jason
Lee wrote in 1834 that The buffaloe [sic] have eaten
nearly all the grass. Near Crowheart in the Wind River

biomass in the semi-arid West.83 As with most ecological

Basin, fur trapper James Clyman wrote about Buffalo

phenomena, explaining the effects of grazing requires

plenty and how a thousand or more bison were killed

an understanding of the interaction of several factors.

on one November day in 1823. In the Green River

Grazing pressure can also cause a decline in biologi-

Valley, in 1811, Wilson Price Hunt wrote, We were

cal crusts on the soil surface of some intermountain

surrounded by mountains in which were disclosed

basins.84 The crusts are growths of cyanobacteria

beautiful green valleys where numerous herds of bison

(former
ly known as blue-green algae), lichens, algae,

graze; which made them much more interesting to us,

mosses, and fungi. They minimize erosion, facilitate

because, for several days, we had not seen a single one

nutrient cycling, and create favorable conditions for

of these animals.87

seedling establishment. Possibly, the crusts were less

As noted, biological soil crusts are easily destroyed

common in Wyoming than in the Great Basin to the

by hooved mammals, whether bison or livestock. Such

west, because of a higher level of herbaceous plant cover

crusts are found in Wyoming, but not as commonly as

that attracted more bison and pronghorn, or because

in the Great Basinpossibly because of livestock graz-

temperatures were too low at the higher elevations

ing since the late 1800s, or possibly because there were

when adequate soil moisture was available. Thus far,

more bison in Wyoming prior to the arrival of livestock.

Sagebrush
Fig. 7.16.The population of feral horses
commonly exceeds the carrying capacity of the rangelands they occupy. Photo
by Ken Driese.

If cheatgrass has not become the widespread problem in

condition in some places. Debate about the comparative

Wyoming that it is in the Great Basin, the explanation

effects of native ungulates and domestic livestock will

may be partially because Wyoming plants evolved with

continue.89

larger herds of bison. If true, native herbaceous plants

Several species of horses evolved in North America,

would have been more competitive and more resistant to

but all became extinct about 10,000 years ago, along

invasive species. As discussed for grasslands, maintain-

with the camel, mastodon, saber-tooth tiger, and

ing competitive stands of native species may be the best

numerous other large mammals (see chapter 2). Some

strategy for excluding undesired plants (see chapter 6).

species survived in Eurasia and were the ancestors of

The management of Wyoming shrublands is more

modern day Equus caballus. Domestication began about

refined today than it was a half-century ago, but the

3000 bc. Horses were reintroduced to North America in

impacts of excessive livestock grazing in those early

1539, when Hernando de Soto and his conquistadores

years probably exist to this day in some areas. Perhaps

brought a herd of 220 to present-day Florida. Mov-

the sagebrush-dominated rangelands most similar to

ing westward was difficult, even with horses, but the

those of the mid-1800s are located some distance from

Spaniards traveled a circuitous route to what are now

frequently used sources of watertoo far away for

Texas and New Mexico. Along the way, some of their

most cattle to useor where fires have not always been

horses wandered off, were traded, or were stolen. Indi-

suppressed, or where the seeds of cheatgrass or other

ans obtained the horse in the 1600s, and by 1750 horses

invasive plants are not yet in the soil. Do such places

were in Canada. The horse caused dramatic changes in

still exist? Debra Donahue, professor of law at the Uni-

the lifestyle of indigenous people, just as they have for

versity of Wyoming, argues that, because of past and

the many people who raise horses today for a multitude

ongoing environmental damage by livestock, there is a

of purposes.90

legal mandate for stopping livestock grazing on public

Feral horses are now common on the shrublands of

lands.88 Doing this even on public land only, she con-

the southwestern quarter of Wyoming and westward

tends, would facilitate the restoration of both the sage-

into the Great Basin states (fig. 7.16). Some, known as

brush ecosystem and nearby riparian zones. Restoration

mustangs, are similar genetically to the horses brought

of sagebrush-dominated shrublands to their prehistoric

to America in the 1500s. With few predators to keep

condition may be expecting too much, but it is plau-

their populations in check, these large herbivores

sible that such initiatives will lead to better rangeland

became, essentially, an invasive species, often forming

127

128 Plains and Intermountain Basins

large herds that graze the rangeland excessively. In 2012

national incentives for energy independence that may

the Bureau of Land Management estimated there were

result in further loss and fragmentation of sagebrush

about 3,500 horses in Wyoming.

habitat. Also, fire commonly has been prescribed for

91

To maintain the feral horse population at appropriate

increasing forage availability for livestock. Indirectly,

levels, horses are rounded up periodically and offered

climate change could become a factor if precipitation

for adoption or for relocation to sanctuariesthe only

patterns change so that deep-soil recharge during spring

option, by federal law, for adjusting horse numbers to

snowmelt is not adequate to support big sagebrush. All

habitat conditions. This method of control has been

such factors interact in troublesome ways, to the point

questioned, because the number of people desiring

that the sagebrush steppe is arguably an endangered

the horses can be insufficient to keep the population

ecosystem.93

in check. Research is under way to develop a practical

Of all the declining sagebrush-dependent species,

method of birth control, but some observers have con-

the greater sage-grouse has received the most attention

cluded that more drastic measures are needed. Despite

(fig. 7.17)in part because it has a curious courtship

the damage they can cause to rangelands, many people

display in the spring. More importantly, the U.S. Fish

enjoy seeing a herd of feral horses loping across the

and Wildlife Service concluded in 2010 that this bird

plains, a vignette of the western frontier.

warrants protection under the Endangered Species Act.


Considering the adverse socioeconomic consequences

Disturbing Trends in Sagebrush-Dependent


Animals: Greater Sage-Grouse

of listing the sage-grouse (because the energy and


agricultural industries would be affected), conserving
the habitat for this species has become a priority for

Considering how much of the West is dominated by

many sectors of society. The Natural Resources Con-

sagebrush, it should not be a surprise that some ani-

servation Service, part of the U.S. Department of Agri

mals require this habitat, most notably, pygmy rabbit;

culture, now has a sage-grouse Initiative; and the state

sagebrush vole; and various birds, including Brewers

of Wyoming has identified core areas that are critical

sparrow, sage sparrow, sage thrasher, and greater sage-

for the maintenance of sage-grouse populations (see

grouse.92 Other species are common in sagebrush eco-

chapter18).94

systems, though they seem to survive in other habitats

There are several species of grouse in North America,

as well, for example, coyote, pronghorn, and mule deer.

each adapted to different ecosystemsan evolutionary

More than 200 species of vertebrates are associated with

process known as adaptive radiation that results in less

big sagebrush in one manner or another.

competition from closely related species. Not surpris-

Since the 1990s, the abundance of sagebrush-

ingly, the large land area occupied by sagebrush pro-

dependent species has been declining because of

vided a niche for one of them, the greater sage-grouse.95

habitat loss caused by various factors, including too-

Through evolution, this bird acquired adaptations that

frequent fires; inappropriate livestock grazing in some

enable it to eat sagebrush leaves, along with insects and

places; and habitat fragmentation stemming from road

forbs. For protection from predators, including ravens,

building, plowing, the construction of homes in rural

coyotes, and eagles, birds of all ages have color pat-

subdivisions, wind farms, and the extraction of oil, gas,

terns that provide camouflage as they hide under the

coal, and uranium. Anything causing a decline in both

shrubs and grasses. Moreover, the sage-grouse are large

shrub and herbaceous plant cover could lead to less

(the largest grouse on the continent), which is adaptive

food availability and less protection from predators for

for the often cold environment they occupy. The birds

sagebrush-dependent species. As the remaining tracts of

range over areas up to 30 miles across during the year,

sagebrush steppe become smaller, the impacts of distur-

but return faithfully to specific places known as leks for

bances have become significantly greater.

courtship and reproduction.96 Breeding adults do not

Reversing such trends is one of the most formid

readily adopt or develop new leks, and they avoid using

able challenges facing land managers today. There are

leks close to roads with considerable traffic at the time

Sagebrush

Fig. 7.17. Male (left) and female (right) greater sage-grouse.


The male is strutting on his lek in early spring. Because of
habitat fragmentation and other factors, the abundance of

of courtship. The birds also avoid leks that have been


compromised by the noise of traffic and drilling.97
Conservation biologists and some land man
agers

this bird has been declining throughout the West during the
past 20 years. In 2012 about 38 percent of the birds lived in
Wyoming. Photos by Mark Gocke.

To further improve sage-grouse habitat, new ap


proaches to livestock grazing are being considered,
with the goal of maintaining adequate grass and forb

have now taken steps where possible to conserve or

cover under the shrubs. This was less of a concern when

improve sage-grouse habitat. The leks are protected

sage-grouse were more abundant and large expanses of

from disturbances associated with new wells, wind

sagebrush steppe without industrial development still

farms, and other developments, and also by keep-

existed. Using small radiotransmitters attached to birds,

ing nearby traffic to a minimum during the breed-

habitat biologists have found that hens in relatively dry

ing season. Keeping tall structures to a minimum is

areas select sagebrush communities with more grass

important, as they provide perches for predatory rap-

cover and litter for nesting and raising their broods.

tors.

New roads are constructed along rights-of-way

Research is under way to determine whether thinning

that avoid critical habitat; the burning of sagebrush

the old shrubs mechanically leads to an increase in

is discouraged or done more judiciously.99 Fires did

nesting habitat as well as the establishment of young

occur historically, probably with minimal effects on

sagebrush.100

98

the sagebrush-dependent animals, but with current

But habitat degradation and road traffic are not the

levels of habitat loss and fragmentation, such events

only factors in play that cause population declines.

can have significant adverse effects in many areas.

Sage-grouse are susceptible to the exotic West Nile

Un
fortunately, as discussed previously, fire manage-

virus, which causes lethal encephalitis in the birds.

ment has been confounded by the invasion of cheat-

Transmitted by mosquitos, this disease was first iden-

grass, which increases fire frequency.

tified as a contributing cause of sage-grouse decline in

129

130 Plains and Intermountain Basins

2003 and was associated with increased mosquito abun-

that could result from warming could cause an unstop-

dance, caused most likely by wastewater ponds created

pable decline in big sagebrush cover.102

during the extraction of coalbed methane.

101

Hundreds

of new ponds were constructed in sage-grouse habitat

Overall, traveling by car or foot though intermountain

in the Powder River Basin alone. Attempts to identify

basins offers views to the horizon of a seemingly nat-

the virus as the cause of population declines have been

ural sagebrush ecosystemmonotonous to some but

confounded by an extended drought that occurred at

inviting to others. The changes over the years are some-

the same time, which could have reduced food avail-

times subtle and not obvious from the road, but aerial

ability. Disease and drought together, along with habi-

views reveal miles of new pipelines, transm ission lines,

tat fragmentation, surely contributed to the alarming

and roads constructed for resource extraction. Recon-

80 percent decline in sage-grouse numbers in the Pow-

ciling the needs of threatened wildlife populations

der River Basin from 2001 to 2012.

with fossil fuel extraction is one of the greatest chal-

The control of road construction, traffic, prescribed

lenges facing western states.103 State governors have

fires, well location, and livestock grazing is feasible,

taken unusual steps in an attempt to reverse the trend

but now the less-controllable effects of climate change

for sage-grouse, as discussed further in chapter 18.

must be considered as well. Warming could benefit

Along with the confounding effects of climate change,

sage-grouse, possibly by favoring the production of the

the issues involved are complex and surely will affect

insects required by chicks. Unfortunately, warming also

the way people benefit from sagebrush-dominated eco-

favors the spread of West Nile virusand the droughts

systems in the future.

Desert Shrublands
and Playas
Chapter 8

Occupying the driest parts of the intermountain

shrubs are Gardners saltbush, winterfat, birdfoot sage-

basins, desert shrublands cover about 8 percent of

wort, shadscale, fourwing saltbush, and greasewood.1

Wyoming. The mean annual precipitation is less than

The dominant plants form a patchy mosaic associated

8 inches, and the generally fine-textured soils tend to

with the amount and timing of water availability, tem-

be saline (see figs. 1.2, 1.5, 3.4, and 8.1). The predomi-

perature at the time water is available, soil texture and

nant plants are different from those in grassland and

salinity, and depth to groundwater (fig. 8.2). Plants

sagebrush ecosystems, though big sagebrush can be

grow slowly and support fewer large herbivores than

found, especially where snow drifting occurs. Typical

in wetter zones.

Fig. 8.1. Desert shrublands


are found where the average annual precipitation is
8 inches or less. Common
plants include Gardners
saltbush, shadscale, fourwing
saltbush, birdfoot sagewort,
bud sagewort, pricklypear cactus, winterfat, and Wyoming
big sagebrush. Greasewood
grows adjacent to playas or
where groundwater seeps to
near the surface. Most of the
plants are shrubs. This area
is in the southern part of
the Great Divide Basin (also
known as the Red Desert),
between Rock Springs and
Rawlins (see fig. 1.2). Elevation about 6,800 feet.

131

132 Plains and Intermountain Basins

High Big sagebrush


Winterfat

Gardners
Shadscale saltbush

are less than in grasslands and sagebrush steppe, primarily because of reduced water availability,3 and perennial
grasses and forbs are less abundant. Fires are infrequent.
Variation in plant distribution can be caused by slight
changes in topography. Playas occur in low depressions,

ARIDITY

where water sometimes accumulates before evaporating,


leaving a dry soil surface that is white with precipitated
salts.4 Plant life is sparse in such areas.
Salt accumulation in desert soils is a formidable problem for the survival of many plants, especially seedlings.
Most notable is that many plants cannot extract water
from soil if the salt (solute) concentration of the soil water
is higher than that of the root cells. Plants known as halo-

Greasewood
Low

phytes have evolved adaptations that solve this problem

Pickleweed
Low

SALINITY

by maintaining cellular concentrations at higher levels

High

Fig. 8.2. Distribution of six desert shrubs in relation to gradients of salinity and aridity. All are found in Wyoming except
pickleweed, the most salt-tolerant species, which is found in
Utah and Nevada. Adapted from West (1988).

than those found in the soil solution. In the leaves of


some halophytes, the salt concentrations promote water
uptake that causes leaf succulence. Greasewood and red
swampfire are examples.5 Other halophytes have evolved
special adaptations for excreting salts (fig. 8.3).6
Halophytes essentially function as pumps, collecting
salts from throughout the rooting zone and depositing

Except for soil salinity, desert shrublands have many

them in a smaller area as leaves fall. Some ecologists

features in common with sagebrush steppe: dominance

have hypothesized that salt deposition under shrubs

of drought-tolerant shrubs with small leaves; exten-

might prevent the establishment of other plants nearby,

sive, sometimes deep root systems for obtaining water

reducing competition for limited water and nutrients

and nutrients; sensitivity to various kinds of herbivores

a form of allelopathy.

and invasive plants; and processes that enhance the

As in grasslands and sagebrush steppe, competition

efficiency of water and nutrient use, such as hydraulic

for resources determines which desert plants and animals

redistribution in the soil and nutrient reabsorption from

manage to survive. This competition is often minimized

leaves prior to leaf fall. Plant cover and annual growth

because of seasonal differences in activity, or because

Fig. 8.3. Scanning electron micrograph


of salt-secreting glands on the leaves of
a salt-tolerant plant in the genus Cheno
podium. The bulbous glands, which are
modified leaf hairs, enlarge until they
burst, after which rainwater washes
the salts onto the soil. Magnification
about 33 times. Microphotograph by
HowardC. Stutz.

Desert Shrublands and Playas

different species tap different parts of the ecosystem for

shrubland, saltgrass meadow, desert grassland, saltbush

the water and nutrients they need (see chapter 6). One

shrubland, and mixed desert shrubland. Greasewood

study focused on how coexisting winterfat and shadscale

shrubland and saltgrass meadow are characteristic of

differed in their adaptations.7 The results showed that

comparatively wet depressions (fig. 8.4), as described

winterfat, a C3 species, is capable of rapid photosynthesis

in chapter 5, but they also are found on some slopes

in the spring and then becomes relatively inactive during

where water seeps to the surface, as sometimes occurs

the rest of the year, when moisture is limiting. In con-

in badlands. The desert grassland community occurs

trast, shadscale, a C4 species, has moderate rates of photo-

where soils are drier and less saline, where drainage is

synthesis throughout the growing season, about half that

not impeded by fine-textured soils, and on the lee sides

of winterfat in the spring but two times faster in the sum-

of slopes where snowdrifts occur. Saltbush shrubland

mer. Both species have a high tolerance for water stress,

is the community with the highest salinity and has an

and both are able to fix carbon at low temperaturesan

abundance of plants in the genus Atriplex. The fifth com-

important adaptation for the cool, short growing season

munity, mixed desert shrubland, is found on soils that

that characterizes northern deserts. Differences between

are intermediate in salinity and have a diversity of shrub

the two species help explain how they coexist. Other des-

species. Sagebrush steppe is sometimes thought of as

ert shrubs have similar adaptations.

desert shrubland, but it occurs at slightly higher eleva-

Sharp spines are another notable adaptation of some

tions or where annual precipitation is somewhat greater.

desert shrubs, especially those that are palatable. This

As described in chapter 7, the sagebrush community has

adaptation minimizes grazing in an environment where

a higher cover of perennial grasses and is likely to burn

the production of plant tissue is slow. Examples are

more frequently than desert shrub vegetation.

greasewood, spiny hopsage, bud sagewort, and shadscale.

Greasewood does not tolerate long periods of

Also, like big sagebrush, some desert shrubs produce both

drought. A deciduous shrub, it commonly grows to

ephemeral and overwintering leaves, improving the effi-

a height of 3 feet or more, and the roots have been

ciency of water use.8 A good example is shadscale.

observed at depths of 10 feet. Because of salt glands


on the leaves, soil salinity under the shrub is typically

The Desert Mosaic

higher than between shrubs.9 Notably, greasewood


seedlings can survive under the parent shrubs where

Five communities can be easily identified in the des-

the salinity is highest, though germination occurs

ert mosaic, based on the dominant plants: greasewood

early in the spring when the soil solution is less saline

Fig. 8.4. Greasewood shrublands in


the Great Divide Basin. Saltgrass and
Gardners saltbush are also common
in this area. The soil surface is often
white, because salts accumulate as water
evaporates. Elevation 6,550 feet. Photo
by Ken Driese.

133

134 Plains and Intermountain Basins


Fig. 8.5. Greasewood is common on
terraces along many ephemeral streams,
with basin big sagebrush and plains silver sagebrush in ravine bottoms, where
salts are washed downstream more
frequently.

because of dilution by spring snowmelt and rains.10

tern is for saline meadows with inland saltgrass and its

Root sprouting also produces new shrubs. At least four

associates to be dotted with low mounds, where much

kinds of greasewood communities have been identi-

taller alkali sacaton grows. The mounds very likely are

fied: greasewood-shadscale, greasewoodbig sagebrush,

the result of silt deposition around the taller grass. Blue

11

greasewood-grass, and a greasewood monoculture.

grama and western wheatgrass can be found with the

Big sagebrush occurs with greasewood only sporadi-

saltgrass in some places, though these species are less

cally, where the soils have a higher infiltration rate and

tolerant of soil salinity.

are less saline.

Inland saltgrass has been the subject of several stud-

Normally greasewood grows at low points in the

ies because of its value as a forage plant. It normally

landscape, but an interesting anomaly sometimes occurs

grows at relatively low levels of salinity.14 Interestingly,

in desert riparian zones, where greasewood shrublands

saltgrass does not concentrate salts in its tissues. Seeds

are found on terraces and basin big sagebrush in ravine

of this grass germinate only when relatively high soil

bottoms, closer to the water table (fig. 8.5). Basin big

moisture coincides with optimal seedbed temperature,

sagebrush grows quite well in the ravine environment,

and when the available soil water is adequate to dilute

probably because of frequent flooding that washes away

the salts. The same might be true for many desert plants.

many of the salts that might otherwise accumulate and

Mixed desert shrubland occurs in dry upland habitats

inhibit the establishment of sagebrush seedlings. Also,

(fig. 8.6). Characteristic shrubs are shadscale, bud sage-

the ravine bottom soils tend to be more permeable than

wort, winterfat, spiny hopsage, and Gardners saltbush

terrace soils, which is favorable for sagebrush.12 Grease-

(table 8.1). Shadscale is more common on dry, some-

wood apparently receives adequate water on the terraces.

what sandier soils; spiny hopsage is common on sandy

Some saline depressions have no shrubs and are bet-

soils where snow accumulation is higher; and winterfat

ter referred to as saltgrass meadows. They are charac-

is more common on fine-textured soils (with more silt

terized by inland saltgrass, alkaligrass, alkali sacaton,

and clay). Big sagebrush can be found in ravines, where

and, in wetter areas, alkali cordgrass. Several halophytic

the soil is less saline and where water infiltration into

members of the goosefoot family are present here as

the soil in the spring is higher (fig 8.7).15 Although big

well. One, red swampfire (Salicornia rubra), is a small

sagebrush is thought to be intolerant of saline or alkaline

succulent plant that turns bright red in the fall and

soils, it sometimes occurs with greasewood.16 This might

seems well adapted to slickspot soilsa name derived

be explained by genetic variation,17 or by the fact that

from the slippery soil that develops after a rain because

plant species composition and growth are a function of

of the soil-particle-dispersing effect of high sodium

topography, aspect, snow accumulation, parent material,

concentrations. Red swampfire commonly forms a

and other factorsnot just soil pH or conductivity. No

band closest to the water, with concentric bands of

single factor explains why plants grow where they do.

saltgrass meadow, greasewood shrubland, and mixed

Desert pavement and coppice dunes are common in

desert shrubland occurring as distance from the pond

mixed desert shrublands. Silt and sand are easily eroded

increases (see chapter 5).13 A common vegetation pat-

by wind, leaving a surface of pebbles. The windblown

Desert Shrublands and Playas


Fig. 8.6. Mixed desert
shrubland on the left, with
Wyoming big sagebrush in
ravines, near Green River.
Elevation 6,500 feet.

particles accumulate around shrubs, forming small

products intended for human consumption.19 Any prac-

dunes (see fig. 7.9). Bud sagewort, galleta, greenmolly

tice that exposes selenium-bearing rock also increases

summercypress, and Gardners saltbush are common on

the density of plants that accumulate selenium. Even-

the drier desert pavement soils, whereas big sagebrush,

tually, as leaves fall and the plants die, this selenium

greasewood, Great Basin wildrye, Indian ricegrass,

is deposited on the soil surface, where it is subject to

spiny hopsage, and thickspike wheatgrass are common

redistribution by erosion. If this happens, toxic concen-

on the dunes. The coppice dunes apparently provide a

trations could develop in places where they would not

relatively moist and less saline environment. Excessive

otherwise be expected.

livestock grazing in deserts can promote the formation

Saltbush shrubland, sometimes called salt desert

of coppice dunes, as described for big sagebrush com-

shrubland, is dominated by Gardners saltbusha short

munities in the previous chapter.

shrub also known as saltsage or Nuttalls saltbush (fig.

Further variation in desert shrublands, and in some

8.8). Associated plants include birdfoot sagewort and

foothill shrublands as well, is caused by seleniuman

bud sagewort, both of which are species of Artemisia

element required for proper animal nutrition, but it can

(see table 8.1). This vegetation type occurs commonly

become lethal as concentrations increase. Some plants,

on fine-textured soils developed from shale or alluvium.

such as woody aster, princesplume, goldenweed, and

In some areas the low shrubs are rare and the com-

several species of milkvetch, accumulate selenium in

munity appears more like desert grassland, with blue

their tissues. The presence of these plants, known as

grama common on somewhat sandier soils and western

selenium indicators, has been used to map selenium-

wheatgrass on fine-textured clay soils. 20 On windswept,

rich bedrock, such as certain shales, siltstones, and

shallow soils with relatively low salinity, the dominant

claystones that are carbonaceous or have high concen-

species may be bluebunch wheatgrass, needle-and-

trations of iron. A concern in western states is that sur-

thread grass, and several species of cushion plants (fig.

face mining and irrigation could increase the amount of

8.9; see table 8.1).

18

land where selenium accumulates to toxic levels in both

Though its total biomass is low, Gardners saltbush is

water and livestock forage, and even meat and plant

valued for the forage it produces on winter ranges, and

135

Table 8.1. Some characteristic plants of desert shrublands in Wyominga

Common name

Latin name

Saltbush
shrubland

Mixed desert Greasewood


shrubland
shrubland

Saltgrass
meadow

Desert
grassland

SHRUBS
Basin big sagebrush

Artemisia tridentata ssp.


tridentata

ravines

ravines

Birdfoot sagebrush

Artemisia pedatifida

Bud sagebrush

Picrothamnus desertorum

Little (alkali) sagebrush

Artemisia arbuscula ssp.


longiloba

Wyoming big sagebrush

Artemisia tridentata ssp.


wyomingensis

Douglas rabbitbrush

Chrysothamnus viscidiflorus

Rubber rabbitbrush

Ericameria nauseosa

Fourwing saltbush

Atriplex canescens

Gardners saltbush

Atriplex gardneri

Shadscale saltbush

Atriplex confertifolia

Greasewood

Sarcobatus vermiculatus

Spineless horsebrush

Tetradymia canescens

Spiny hopsage

Grayia spinosa

Winterfat

Krascheninnikovia lanata

Alkali sacaton

Sporobolus airoides

Blue grama

Bouteloua gracilis

Bluebunch wheatgrass

Pseudoroegneria spicata

Thickspike wheatgrass

Elymus lanceolatus

Western wheatgrass

Pascopyrum smithii

Bottlebrush squirreltail

Elymus elymoides

Foxtail barley

Hordeum jubatum

Indian ricegrass

Achnatherum hymenoides

Needle-and-thread grass

Hesperostipa comata

Nuttalls alkaligrass

Puccinellia nuttalliana

Salina wildrye

Leymus salinus ssp. salinus

Saltgrass

Distichlis spicata

Sandberg bluegrass

Poa secunda

Goosefoot

Chenopodium spp.

Green molly

Bassia americana

Hookers sandwort

Arenaria hookeri

Nuttalls povertyweed

Monolepis nuttalliana

Prairie (fringed) sagewort

Artemisia frigida

GRASSES

FORBS

Desert Shrublands and Playas

Saltbush
shrubland

Mixed desert Greasewood


shrubland
shrubland

Saltgrass
meadow

Desert
grassland

Suaeda spp.

Phlox hoodii

Common name

Latin name

Pepperweed

Lepidium densiflorum

Pricklypear cactus

Opuntia polyacantha

Red swampfire

Salicornia rubra

Saltlover (halogeton)

Halogeton glomeratus

Scarlet globemallow

Sphaeralcea coccinea

Seablight
Spiny (Hoods) phlox
Tanseyleaf tanseyaster

Machaeranthera tanacetifolia

Textile onion

Allium textile

Yellow spiderflower
(beeplant)

Cleome lutea

A dash indicates that the plant is absent or uncommon.

the seed is sometimes planted for reclamation purposes,

was found to be 810 inches in central Washington,

especially when the topsoil has high pH and salinity.

of which 1831 percent was transpired.24 As might be

Its roots spread laterally to a distance of 6 feet, with

expected, greasewood height, canopy cover, and total

taproots penetrating to 3 feet or more. Root sprouting

leaf area were highest when groundwater was near the

or layering is the major form of reproduction once the

surface. One study found that greasewood caused less

plants are established. These small shrubs are thought


to live more than 50 years.21

30

is winterfat, so named because some ranchers observed


that their livestock could gain weight by eating this plant
when it was dormant.22 In addition to sheep and cattle,
winterfat is eaten by pronghorn, elk, deer, and rabbits.
It also is one of the most useful shrubs for reseeding
disturbed rangelands where annual precipitation is less
than 10 inches. As with all desert plants, new seedlings
become established only when favorable moisture and
temperature conditions coincide for a long enough time.

The Desert Shrubland Ecosystem


With regard to total annual plant growth, cold-desert
ecosystems with different dominant species have been
ranked from high to low in this order: big sagebrush,
Gardners saltbush, greasewood, shadscale, spiny hopsage, and winterfat.23 The most productive shrublands
undoubtedly have the highest water consumption.
Annual evapotranspiration from a stand of greasewood

WATER INFILTRATION (cm)

Another highly valued forage for livestock and wildlife

Big sagebrush steppe

25
20
15

Grassland

10
5
0

Desert shrubland
0 10

30 45 60

90

120

MINUTES

180

Fig. 8.7. Desert shrublands often have finer-textured soils than


are found in grasslands or big sagebrushdominated shrublands, which accounts for much slower water infiltration (see
also fig. 3.8). In arid regions, less water is available for plant
growth in soils with slow infiltration rates. Adapted from
Nichols (1964).

137

138 Plains and Intermountain Basins

Fig. 8.8. (above) Saltbush desert shrubland northwest of Rock


Springs. The dominant shrubs are Gardners saltbush, bud
sagewort, and birdfoot sagewort. The treeless escarpment is
formed from late Cretaceous sandstones and shales. Elevation
6,500 feet.
Fig. 8.9. (left) Windswept desert shrubland on a rim of the
Great Divide Basin, with cushion plants adapted for conserving water and heat. Common plants in this community are
Hookers sandwort, squarestem phlox, matted buckwheat,
stemless goldenweed, and stemless four-nerve daisy. Similar
growth forms but different species are found in alpine tundra.

soil water drawdown when leafless than did a stand of


evergreen big sagebrush.25 Generally, invasive cheatgrass grows more abundantly in relatively moist stands,
such as where greasewood occurs (see chapter 7 for a
discussion of cheatgrass). Research in Utah indicates
that cheatgrass can invade desert shrublands during
years of above-average rainfall.26
Other

than

greasewood-dominated

shrublands,

which occur where water is more readily available, desert ecosystems are characterized by potential evapotranspiration that greatly exceeds annual precipitation
and the water that might be gained by drainage from
surrounding slopes. Consequently, plants are frequently

Desert Shrublands and Playas

under water stress. Infiltration rates vary considerably


and appear to influence the distribution of different
plants (see fig. 8.7). Nutrient leaching is rare, but low
levels of nitrogen can limit plant growth.27 Nitrogen fixation results from species of cyanobacteria, bacteria, or
lichens found in biological soil crusts, but atmospheric
inputs of nitrogen may be more important.28 Nutrient
losses are of little consequence, primarily because water
is usually the primary limiting factor, but also because
most nitrogen requirements for new plant growth probably are met by mineralization of soil organic matter
and reabsorption from senescing leaves. As in grasslands and other shrublands, herbivores and detritivores
increase the rate of nutrient cycling (see chapter 6), and
the presence of shrubs creates islands of fertility (see
chapter 7).
Except for the effects of livestock grazing, relatively
little is known about plant-animal interactions in desert shrublands, compared to grasslands and sagebrushdominated shrublands. Jackrabbits are thought to prefer
winterfat for food but use nearby patches of big sagebrush for hiding cover (fig. 8.10). In one study, winter-

Fig. 8.10. White-tailed jackrabbit in the Red Desert, near


Adobe Town. Photo by Ken Driese.

fat nearest the sagebrush was subject to heavier grazing


by the hares, which in turn facilitated the expansion
of big sagebrush. In this way the jackrabbits influenced

phytic plant biomass? In the case of greasewood shrub-

the vegetation mosaic, at least during years when their

lands, with finer-textured soils that are wet and possibly

numbers are high. Some investigators believe that jack-

anaerobic during much of the year, is most herbivory

rabbits in large numbers can graze rangelands more

still belowground, as in grasslands and some shrub-

heavily than livestock or big game.29

lands? How do high concentrations of selenium affect

In addition to jackrabbits, Townsends ground squir-

microbial activity?

rels can be abundant in desert shrublands during some


years, providing prey for raptors.30 In southern Idaho,
the rodent populations drop significantly when desert

Disturbances and Succession

shrublands are converted to communities dominated

The major disturbances in desert shrublands are

by annuals, such as cheatgrass.

drought and extraordinarily heavy herbivory, whether

Little research has been done on the hydrology,

from grasshoppers, Mormon crickets, bison, or live-

energy flow, and nutrient cycling of Wyoming desert

stock. Fires are less frequent in more arid environments

shrublands, but some interesting opportunities exist.

because of insufficient fuel, but they surely occur in

For example, just as it is helpful to consider the effects

greasewood and mixed desert shrublands during years

that the presence of big sagebrush has on ecosystem

when fuel accumulates, especially after periods of rela-

properties (see chapter 7), it also is interesting to con-

tively high rainfall that coincide with light grazing, or

sider the effects of salinity and selenium. Does nitrogen

where cheatgrass has invaded (see chapter 7). Harvester

become more or less available as salinity increases? Are

ants are abundant in some saltbush shrublands, creat-

soils high in sodium more susceptible to erosion? How

ing patchy disturbances,31 and sometimes two years or

do patterns of herbivory and decomposition change, if

more of drought will cause portions of a shrub to die,

at all, because of higher salt concentrations in the halo-

if not the entire plant.32 Shadscale is apparently very

139

140 Plains and Intermountain Basins

sensitive to drought, with dramatic annual variations

In addition to the loss of soil crusts, introduced

in plant cover associated with spring precipitation and

annual plants have greatly affected some desert shrub-

the abundance of a scale insect that feeds on roots.33

lands, just as they have some shrublands dominated by

Drought is not the only explanation offered for des-

sagebrush (see fig. 7.13). Three common weeds are cheat-

ert shrub mortality. Between 1977 and 1986 the shrubs

grass, halogeton, and Russian thistle.40 Of the three,

over large areas in the Great Basin of Utah died, appar-

cheatgrass is by far the most invasive and has become

ently stemming from four factors triggered by several

a major problem in western intermountain basins (see

years of unusually high precipitation: (1) frequent

chapter 7). Some varieties of the species are adapted to

periods when the soil was anaerobic, (2) increased soil

tolerate saline soils and can invade even with little or no

salinity, (3) increased susceptibility of shrubs to soil-

livestock grazing.41 As in sagebrush, cheatgrass typically

borne diseases, and (4) the loss of shrub tolerance to

grows so densely in favorable years that fires burn in

the droughts that occurred after the wet period ended. 34

places where previously they could not because of in

Thus far, a similar dieback of desert shrubs has not been

sufficient fuel. And the fires destroy the patches of soil

reported for Wyoming.

crust that might still persist. Arguably, cheatgrass is the

Excessive livestock grazing has been a common

most severe problem facing desert shrubland managers

disturbance in most desert shrublands. 35 Cattle drives

in Utah, Nevada, and southern Idaho, and with climate

first started on the Great Plains, where the rangelands

change and ongoing evolution, it could become a severe

were adapted to the grazing of large mammals because

problem in Wyoming as well. Douglas Shinneman and

of large numbers of bison there. Moving westward, the

William Baker compared many sites in the Great Basin

cattle foraged in big sagebrush steppe, which also had

states,42 finding more rapid cheatgrass invasion where

considerable grass cover and a long history of grazing by

soil crust cover had declinedanother indication that

bison. Sometimes the concentrations of cows and sheep

managers must consider microbial organisms along

were much higher than livestock managers would allow

with large animals, fire, and the dominant plants as

today, and without careful herding, the same numbers

they strive to achieve sustainable land management.

of animals easily wandered into the adjacent desert

Though not yet reported for Wyoming, an unusual

shrublands, causing dramatic changes.36 Some of the

effect of cheatgrass invasion was to convert a shrubland

palatable plants declined with this added grazing pres-

in eastern Washington dominated by big sagebrush to

sure, while others increased. 37 Recovery after reducing

one dominated by greasewood. The following mecha-

the numbers of livestock was very slow because of the

nism was proposed by William H. Rickard: (1) a high

arid environment.

cover of cheatgrass led to the interception and use

One of the first adverse effects of livestock graz-

of more of the annual precipitation, resulting in in

ing was probably the degradation of delicate biologi-

adequate percolation of water to the depths required by

cal soil crusts, where they occurred, by trampling (as

big sagebrush; (2) fire frequency was increased by the

discussed in chapter 7). 38 When this happens, soil ero-

flammable cheatgrass, but greasewood (already present

sion increases, nitrogen fixation may be reduced in

in the community) persisted because it could sprout,

some years, and favorable sites for seed germination

unlike big sagebrush; and (3) greasewood caused an

and seedling establishment can be lost. Considerable

increase in surface soil salinity, which big sagebrush

research has been done on the importance of the crusts

seedlings could not tolerate.43 Similar changes could

for desert shrubland ecosystems in the Great Basin, but

occur in Wyoming if cheatgrass were to become abun-

their importance in Wyoming desert shrublands is still

dant where big sagebrush and greasewood coexist.

uncertain. Today they are most easily found in the Bighorn Basin, in places where grazing has been minimal.
Were the crusts less common in Wyoming to begin

Scenario for the Future

with, because of a higher cover of perennial grasses

Extrapolating from what is known about desert eco

between the shrubs and heavier grazing by bison? An

systems, a warmer climate would increase potential

answer is not possible at this time.39

evapotranspiration and rates of soil drying. There could

Desert Shrublands and Playas

be occasional periods of high precipitation, but the dry

Whether disturbed by excessive livestock grazing or

periods are likely to be longer than during the past cen-

industrial developments, the restoration of desert eco-

tury, and more of the annual precipitation will come as

systems is difficult. Even under experimental conditions,

rain rather than snow. Big sagebrush in Wyoming and

new seedlings typically do not survive, and the devel-

the region will become less common if deep soil water

opment of new soil crusts takes 10 years or more. Soil

recharge is inadequate. Most likely, desert plants will

erosion occurs with less plant cover, leading to the for-

expand to occupy some land now dominated by sage-

mation of desert pavement and coppice dunes. Warmer

brush. Other changes in plant species composition will

temperatures will make reclamation even more difficult

occur as well. For example, already endangered species,

and most likely will cause a decline in the amount of

such as the greater sage-grouse, will lose some of their

soil organic matter, making the soils still more erodible

habitat as big sagebrush declines, and cheatgrass will

and a source of atmospheric carbon dioxidethereby

become more common and will occupy a larger area,

negating attempts to sequester carbon in terrestrial

very likely surviving the droughts (because it grows pri-

ecosystems. The desert shrublands of the twenty-first

marily in late fall and early spring, when sufficient mois-

centuryand the way they are usedare likely to be

ture probably will be available). The increased biomass

considerably different than during the twentieth cen-

of cheatgrass will probably increase fire frequency, caus-

tury, primarily because of mistaken assumptions about

ing further declines in big sagebrush. Carrying capacity

livestock management in the late 1800s, the invasion of

for big game and livestock in the intermountain basins

cheatgrass shortly thereafter, and the climate changes

will then decline.

now under way.

141

Sand Dunes,
Badlands, Mud Springs,
and Mima Mounds
Chapter 9

Sand Dunes
Stretching across the intermountain basins of central
Wyoming are elongated mosaics of active and stabilized
dunes (fig. 9.1). Sand dunes create a dramatically different environment in the semi-arid lowlands, primarily
because they provide a better water supply for plants,
due to the inverse texture effect (explained in chapter 3)
and because the shifting sand creates special problems
for plant establishment and growth. Consequently, the
flora of the dunes is considerably different from that of
the surrounding shrublands. Contributing to the biological diversity in some areas are small ponds that provide a source of water during the dry summer for plants
and animals that might not otherwise be able to survive
there.
Dune location depends on wind direction, a source
of sand, and barriers to sand movement. For example,
large deposits of sand are found on the westerly, windward foothills of the Ferris Mountains north of Rawlins,
the Laramie Mountains near Laramie, the Medicine Bow
Mountains in North Park, Colorado (east of Walden),
and the Sangre de Cristo Mountains (Great Sand Dunes
National Park) farther south in Colorado. Elongated
dunes are found in wind corridors where sand has been
funneled by the topography, such as the Killpecker
Dunes northeast of Rock Springs, and the dune fields
west and east of Casper.
The origin of such large volumes of sand is one of
the first questions asked. The answer lies in the physical
weathering of various kinds of rocks over long periods
142

and the power of water and wind to sort and move the
resulting particles. The sand grains (0.052.00 mm in
diameter) are gradually separated from coarser gravel
and much of the finer silt and clay. The sand then accumulates along creeks and rivers in the intermountain
basins. For example, most sand in the Killpecker Dunes
originated from the appropriately named Big Sandy
and Little Sandy creeks that flow from the Wind River
Mountains.1 These creeks were much larger when the
glaciers were melting, bringing large volumes of sand to
the area. Subsequently, westerly winds in the Holocene
blew much of this sand toward the east, across the Great
Divide Basin to the Ferris and Seminoe mountains.
Smaller silt and clay particles, known as loess, were
dispersed farther eastward onto the Great Plains. The
presence of different-aged sand grains in now-dormant
parts of the Killpecker Dunes suggests that sand movement occurred in this area at various times over the past
20,000 years, probably most often during dry periods,
when the sand was more easily blown out of the creek
beds.2
The kinds of plants growing on dunes depend on
the degree of sand stabilization as well as on temperature, moisture, and the amount of organic matter
in the sand. Only a few species can survive on shifting dunes in Wyoming. They include blowout grass,
sand lovegrass, prairie sandreed, sandhill muhly, rusty
lupine, blowout penstemon, alkali wildrye, Indian
ricegrass, and scurfpea. 3 The coarser soils of dune tops
typically have different species than the finer-textured

Sand Dunes, Badlands, Mud Springs, and Mima Mounds 143


Fig. 9.1. Location of the primary
sand dunes and badlands in
Wyoming. Some dunes shown
on this map are stabilized by
vegetation. Cartography by Ken
Driese.

soil between dunes.4 Alkali cordgrass is common where

Notably, the tallest, most vigorous stands of basin big

wetlands form between the dunes. Missing from Wyo-

sagebrush in the region occur on stabilized dunes, such

ming dunes is American beachgrass, commonly found

as some parts of the Killpecker and Ferris dunes where

on inland dunes to the east and along the Atlantic and

these shrubs grow to 6 feet tall or more. The tallest sage-

Pacific coasts, and yellow wildrye, a common colonizer

brush is typically on the leeward sides of dunes, where

in the dunes of southeastern Idaho.5

snow accumulates. Such sagebrush woodlands provide

Succession in dune ecosystems is primarily a process

important cover for the elk that live in the area.

of soil stabilization (figs. 9.29.4). The continued growth

However, stabilized dunes are subject to the same

of pioneer species gradually leads to more organic

kinds of periodic disturbances important elsewhere

matter in the sand and more soil binding by the root

drought, fire,7 burrowing, heavy grazing, and human

systems, which include the fungal filaments of mycor-

traffic. Such disturbances lead to a decline in plant

rhizae. The mere presence of the plants causes reduc-

vigor, subjecting small areas of the dune to rapid ero-

tions in wind velocity, diminishing the extent of sand

sion. The resulting blowout, created in a few days or

movement. This stabilizing process continues as plant

weeks, is eventually stabilized again through the per-

cover increases; eventually the entire dune is covered

sistence and expansion of the pioneer species. Signifi-

with plants. During stabilization, many other grasses,

cantly, the suppression of fires is thought to have greatly

forbs, and shrubs become common, including antelope

increased the proportion of some dune fields that are

bitterbrush, big sagebrush, rabbitbrush, silver sage-

now stabilized, such as in the Nebraska Sand Hills.8

brush, spiny hopsage, and others. Some plants create

Destabilization occurs with drought, heavy grazing, or

favorable microenvironments for others that could not

the recreational use of all-terrain vehicles.

grow on the dune initially. The pioneer species become


6

less common, but they persist in the community.

Compared with surrounding grasslands and shrublands, some parts of the dune environment are more

Fig. 9.2. (above) Killpecker Dunes,


located northeast of Rock Springsone
of the largest active dune fields in North
America. Dune colonizers in the foreground include Indian ricegrass, alkali
wildrye, slimflower scurfpea, veiny
dock, rusty lupine, skeletonplant, basin
big sagebrush, rabbitbrush, and spiny
hopsage. Elevation 7,000 feet. Photo by
Ken Driese.
Fig. 9.3. (left) Early stages of stabilization in the Killpecker Dunes, caused
here by veiny dock (Rumex venosus).
Aswith other dune-colonizing species,
such as slimflower scurfpea and alkali
wildrye, this plant spreads by rhizomes.
Photo by Ken Driese.

Sand Dunes, Badlands, Mud Springs, and Mima Mounds 145

attracted to the dunes for this reason, including deer, elk,


waterfowl, and wild horses. The ponds also provide habitat for various aquatic organisms, including salamanders,
spadefoot toads, and freshwater shrimp. Because dunes
are an oasis in the semi-arid basins and plains, livestock
may concentrate there, creating management problems
similar to those encountered in riparian landscapes.
Indigenous people were also attracted to the dunes,
undoubtedly because of water and food availability.11
Artifacts, estimated at more than 10,000 years old, have
been found in the Killpecker Dunes and elsewhere in the
region, commonly in association with bison bones.
Although sand dunes are favorable for the growth of
established plants, shifting sands provide a poor substrate for seedling establishment. Moreover, established
plants are easily buried. Pioneer dune plants (such as
blowout grass and scurfpea) survive because their rhizomes grow up or down quickly enough to maintain
appropriate depths for the species. Aside from that, root
growth occurs in the direction of moisture availability.
As important, the rhizomes have considerable energy
stored in the form of carbohydrates, which enables the
growth of numerous shoots when the buds begin to
grow, all characteristics that diminish the probability
Fig. 9.4. Slimflower scurfpea. Note the rhizome and prominent taproots that have been exposed by wind erosion.

that the entire plant will be buried to excessive depths.


Roots also grow from the buried vertical stems of some
plants (adventitious roots), even on seedlings.12
Dune plants encounter other problems as well, such

favorable for the survival of many plants and animals.

as sandblasting of their leaves and stems. Not surpris-

The annual production of new plant biomass can be

ingly, most of them are herbaceous, with shoots that

more than double that of the adjacent sagebrush steppe.9

require maintenance for only a few months. Also, some

Even with limited precipitation, the rapid infiltration

species produce waxy coatings that reduce abrasion by

of water through the coarse sand provides good water

blowing sand. Unlike in the coppice dunes discussed

storage, and commonly snowdrifts on the lee sides of

in chapter 7, woody plants invade only after the dunes

dune crests are buried by drifting sand in the spring. It

have become partially stabilized and abrasion is less

is common to find buried snow in midsummer, which

severe. Sand abrasion is not all bad for the plants, as

provides a source of water for ponds and perennial

some of them have seeds that will not germinate until

streams as it melts.10 The landscape pattern is diversified

seed coats are abraded. Other adaptations contribute

further by the shrubs and small trees that occur along

to water-use efficiency in this environment, where the

such streams, for example, basin big sagebrush, com-

temperature at the soil surface can be high. Many dune

mon juniper, currant, Oregon-grape, Rocky Mountain

plants are water-efficient C 4 species, and a large propor-

juniper, water birch, Woods rose, and willow.

tion have both shallow and deep roots.13

Rapid infiltration reduces the time available for evap-

Extensive root systems with mycorrhizae are impor-

oration. This, along with buried snow that melts slowly

tant for obtaining nutrients, as sand is an infertile sub-

during the summer, contributes to the formation of small

strate. Also, two common dune colonizers in Wyoming

ponds between the dunes (fig. 9.5). Many animals are

are nitrogen-fixing legumes (scurfpea and lupine).14 As

146 Plains and Intermountain Basins


Fig. 9.5. Small ponds sometimes form
between dunes, primarily because of
snowdrifts that melt slowly after being
buried by drifting sand. The colonizing plants in this photo, taken in the
Killpecker Dunes, are alkali wildrye and
slimflower scurfpea. Photo by Bonnie
Heidel.

the plant tissues grow and die, soil organic matter gradu-

turbance, however, the dunes eventually become sta-

ally accumulates. Some of it decomposes, providing

bilized and highly productivea classic example of

nutrients that, combined with water availability, create

natural succession. The dunes of central Nebraska have

one of the best environments for plant growth in the

been one of the most productive grasslands on the

intermountain basins. That may be short-lived, however,

Great Plains, providing forage even during extended

as the next blowoutwhether caused by an extended

droughts, such as occurred in the 1930s.

drought, fire, or other disturbancemay bring the

Most sand dunes in Wyoming occur at higher eleva-

organic matter to the surface, where it is widely dispersed

tions than in adjacent states, and consequently they

by the wind. Alternatively, the blowing sand may cover

have a relatively cold climate and short growing sea-

well-developed soils. Known as paleosols, such buried

son. With current predictions for increased warming

soils are useful for interpreting climate history.

and drought, reactivation of dunes over large areas

A newly activated dune generally has a deficiency

seems likely and may have already begun. Numerous

of nutrients on the surface. Without drought or dis

studies have shown that drought causes a rapid decline

Sand Dunes, Badlands, Mud Springs, and Mima Mounds 147

in plant cover on stabilized dunes, which leads to the

ming, fossils of crocodiles, primates, and tapirs date

dune becoming reactivated.15 The effect of drought is

back to the Cretaceous and early Tertiary periods, indi-

hastened if fires, intensive grazing, disruption by off-

cating a subtropical climate. Some badlands are now

road vehicles, or other disturbances occur at about the

protected because of their value for paleontological

same time.16 Thus far, few invasive plants have been

studies and the number of people fascinated by such

observed in Wyoming dunes, but that is not true at

places. The most notable example is Badlands National

lower elevations.

Park in western South Dakota.

17

Plants are sparse in badland landscapes. The few

Badlands

that can be found also occur in adjacent grasslands


and shrublands, such as big sagebrush, bottlebrush

Some parts of the lowlands, known as badlands or breaks,

squirreltail, Gardners saltbush, juniper, shadscale, sil-

are nearly devoid of vegetation because of naturally high

ver sagebrush, tufted evening primrose, and western

rates of erosion associated with retreating escarpments

wheatgrass. Halophytes are common. In ravines or near

(figs. 9.6 and 9.7). Such areasfor example, Adobe Town,

seeps, basin big sagebrush, basin wildrye, greasewood,

Grizzly Buttes, Hells Half Acre, Honeycomb Buttes, the

narrowleaf cottonwood, and species of willow may be

Powder River Breaks, and Wind River Badlandsare some

present. The amount of plant cover varies considerably

of the most fascinating landscapes in the region. Badland

with elevation, substrate, precipitation, and topographic

topography, sometimes angular and sometimes rounded,

position. On steep slopes, erosion is a nearly continuous

occurs in arid climates on shales and mudstones. The

process, which prevents plant establishment.

clay-rich soils expand and contract with wetting and

Because of the abrupt, finely dissected topography,

dryingconditions that usually prevent plant establish-

the fauna of badlands can be more unusual than the

ment and facilitate erosion. Flash floods are common

flora. Hawks, eagles, swallows, and swifts find suitable

because of slow infiltration during thunderstorms.

nesting sites on exposed cliffs, and larger animals use

Because badlands are formed by rapid erosion, they

the badlands year-round, finding shelter on leeward

are favorite destinations for fossil hunters. Each year

slopes on cold, windy days. The vegetation mosaic is

new specimens are exposed on the surface. In Wyo-

unusual as well. In the Powder River Breaks of north-

Fig. 9.6. Honeycomb Buttes in


the Jack Morrow Hills northeast of Rock Springs, one of
numerous tracts of badlands
found in Wyoming. Here the
rocks are shales, claystones,
and sandstones of the Green
River and Wasatch formations
(Eocene). Elevation 6,888 feet.
Photo by Ken Driese.

148 Plains and Intermountain Basins


Fig. 9.7. Skull Rim in a desert
wildland known as Adobe
Town, part of the Washakie
Basin southeast of Rock
Springs and southwest of
Rawlins. The badlands in this
area are carved from claystones and sandstones of the
Washakie Formation (Eocene).
Photo by Ken Driese.

ern Wyoming and southeast Montana, seven types of

surrounded by mudflats or playas with only a few plants

shrublands and woodlands have been identified, based

(fig. 9.8). The cones seem to provide better conditions

on their dominant plants: shadscalebig sagebrush,

for plant growth than the flats do. At one time, a slurry

big sagebrushshadscalewestern wheatgrass, big sage-

of mud oozed from their summits and then dried.19

brushwestern wheatgrass, greasewood, skunkbush

The geologist and explorer Ferdinand Hayden was

sumacwestern wheatgrass, Rocky Mountain juniper

the first scientist to discover these unusual landforms

western wheatgrass, and ponderosa pinejuniper. In

when he visited the area in 1877. According to his

addition, on mesa tops capped by erosion-resistant

report, the mounds at that time had pools of muddy

scoria, there are small patches of mixed-grass prairie.

water at the top. Observing that a bubble of gas periodi-

Greasewood is found near seeps on some slopes. Plant

cally rose to the surface of the pools, he and his men

cover is sometimes contoured in narrow bands that par-

conducted an experiment:

allel beds of gray clay, buff-colored silt, and lignitean


indication of strong edaphic control.18
Ponderosa and limber pine are found in badlands,
along with Utah and Rocky Mountain juniper, but only
on escarpments that are comparatively resistant to erosion. Elsewhere erosion occurs before mature trees
have time to develop. Landslides can cause significant
changes.

A rifle-ball shot down vertically into one of the


openings produced a sudden eruption of the whole
mass. Water and mud were thrown to a height of
about 10 feet, covering the luckless experimenter
from head to foot. From a safer distance the trial was
several times repeated and almost always followed
by the same result. . . . Crude as this test may be, it
shows the presence of gas at some depth, held there
under mechanical pressure. 20

Mud Springs

Today, the mud springs appear to be dormant. A

One of the most unusual features of the Great Divide

possible cause is the pumping of groundwater from the

Basin is the small area of mud springs found just east of

Great Divide Basin for livestock and industry, which

Circle Bar Lake (one of the Chain Lakes north of Wam-

could be relieving some of the water pressure that

sutter). These conical mounds are 315 feet high and are

was important for the creation of these springs. How-

Sand Dunes, Badlands, Mud Springs, and Mima Mounds 149


Fig. 9.8. Mud springs in the Chain
Lakes area, located in the lowest part
of the Great Divide Basin, are currently
dormant. Plants include greasewood
and shadscale. A playa with little or
no vegetation surrounds the conical
mounds. Elevation 6,500 feet.

ever, it is possible that the mounds are rejuvenated by

Mima mounds raise two questions: What is the

periodic rather than continuous eruptions. Hayden

effect of the mounds on plant and animal distribu-

observed deep pools of muddy water in the area near

tion, and how did they originate? The effect on plants

the mounds, covered by a thick layer of crusty silt that

can be subtle, with no obvious differences between

gave the appearance of solid ground. He wrote, Innu-

mound and intermound vegetation. Elsewhere, the

merable bones of animals, who here sought to quench

vegetation on the mounds is taller or composed of

their thirst, provide the treacherous character of the

different species. For example, the mounds may have

soil. A nearby pond was christened Death Lake by

more big sagebrush and higher plant growth rates than

the explorers.

between mounds. 22 Some investigators have noted


that burrowing mammals, such as pocket gophers and

Mima Mounds

ground squirrels, are more common on the mounds.


Through their burrowing, the mammals create soils

A peculiar feature of some lowland grasslands and

that have a reduced bulk density (weight per unit vol-

shrublands in Wyoming are hundreds of closely spaced

ume), enhanced infiltration rate, and improved nutri-

mounds that are 1224 feet in diameter and usually less

ent availability. Vegetation differences can be expected

than 2 feet high (figs. 9.9 and 9.10). They are common on

when mound and intermound soils are different, but

the east slope of the Laramie Mountains (west of Chey-

whether the differences are due to the soil or associated

enne), in the Laramie Basin southwest of Laramie, and

animals is still unknown. Unfortunately, cheatgrass

in the Shirley Basin north of Medicine Bow (all at low

has been found on the tops of some mima mounds in

elevations). Equally puzzling mounds occur in numerous

the Laramie Basin.23

other places in North America and in Argentina, Kenya,

The more perplexing question about mima mounds

and South Africa.21 Known as prairie mounds, pimple

pertains to their origin. There are various hypotheses

mounds, and biscuit land, they are most commonly

that involve wind deposition (coppice dunes), erosion,

referred to as mima or mima-like mounds, because they

frost heaving, groundwater flows, burrowing animals,

were first described on the Mima Prairie in western

and even seismic activity. The erosion hypothesis sug-

Washington.

gests that the tops of mounds were once the location of

150 Plains and Intermountain Basins

Fig. 9.9. (above) Mima mounds in the Laramie Basin. The vegetation is high-elevation mixed-grass prairie, dominated by
junegrass, Sandberg bluegrass, Indian ricegrass, needle-andthread grass, blue grama, fringed sagewort, and Hoods phlox.
Fig. 9.10. (left) Aerial view of mima mounds in the Laramie
Basin. In this area the mounds are partially on the broad
floodplain of the Big Laramie River (see chapter 17). Elevation
7,250 feet.

a stabilizing tree or tall shrub, with the soil in between


more susceptible to erosion. After the trees or shrubs
died, the mound persisted, though possibly with some
smoothing by wind. Indeed, the mound distribution
patterns are often similar to tree distribution patterns
in woodlands and savannas, but there is little or no supporting evidence for this explanation.
The frost hypothesis has received more attention than
the deposition and erosion hypotheses in Wyoming.
From the air, the mound pattern gives the impression of
sorted polygons that are common in the arctic tundra
(see fig. 9.10), an observation that suggests the mounds
were formed about 20,000 years ago when there were
glaciers in the mountains and the climate was colder.
Indeed, Wyoming intermountain basins at that time had
permafrost with relatively little snow cover.24 In such
environments, soil cracks 3 feet deep or more were cre-

Sand Dunes, Badlands, Mud Springs, and Mima Mounds 151

ated by contraction caused by either freezing or desicca-

published in 1954, R. J. Arkley and H. C. Brown proposed

tion. These cracks filled with windborne soil particles,

the following explanation:

and today their approximate form and depth are apparent when the soil profile is exposed. Depending on the
cause of the cracks, the distinct patterns in the soil profile are referred to as fossil ice wedges (see fig. 2.8) or sand
wedge relics, the latter having been formed in desiccation cracks.
University of Wyoming soil scientists Lowell Spackman and Larry Munn dug trenches through mima
mounds in the Laramie Basin and determined that fossil ice wedges were a key factor in mound formation. 25
They concluded that the mounds could be attributed
to cryostatic pressure created from water entrapped
between a layer of permafrost or bedrock and a downward-freezing frost layer from the surface. Pressure was
ultimately released through planes of weakness created

The method by which the [pocket] gopher accumulates soil into a mound is explained by his tendency
to place his nest in a well-drained spot where the
soil is deepest; thus generation after generation of
gophers may keep building nests near the crest of
any high spot in the land surface, or over a window
in a hardpan. When a gopher is tunneling, he moves
the soil beneath his body, and forces it backward to
a surface opening already established. . . . Thus, over
a long period of time, the gopher, by digging outward from his nest, tends to move the soil toward it.
. . . The mound rises very gradually over a considerable period of time and perhaps many generations of
gopher occupation.28

by large fossil permafrost sand wedges. They found no

Others, primarily ecologist George Cox and his asso-

evidence for the burrowing rodent hypothesis. Previous

ciates, concluded that burrowing animals are the uni-

research had suggested that the mima mounds in Wash-

versal cause of mima mounds everywhere.29 Cox wrote,

ington were also formed in association with permafrost

Fossorial rodents are the builders of Mima mounds,

and ice wedges.26

the largest and most widespread landscape features pro-

Working in the same area, soil scientist Richard

duced by any mammal other than man. Cox reviewed

Reider and his associates proposed a mound-forming

studies indicating that, in addition to pocket gophers,

mechanism that involved low-pressure groundwater

mound building is caused by ground squirrels; badgers;

flow.27 They dug trenches through the mounds as well,

toads; fire ants; and, in Kenya, mole rats.

finding that all of them had (1) shallow impermeable


bedrock, (2) a thin layer of alluvial gravel on top of the

The rationale of Cox, suggesting that burrowing created

bedrock, and (3) well-developed soil at the ground sur-

all mounds worldwide, still needs to be reconciled with

face. They also observed that the mound soil was lens-

the evidence obtained supporting other hypotheses in

shaped in cross section and that it had concentrations

the Laramie Basin. A new dimension to the debate was

of sodium in patterns suggesting a swirling ground

added by geologist Andrew Berg, who concluded that

water flow. Reider and his associates concluded that the

most mima mounds are the result of seismic vibrations

explanation probably lies in blending their explanation

in Earths crust in areas with unconsolidated fine sedi-

with Spackman and Munns cryostatic hypothesis, but

ments on a relatively rigid, flat substratum. This led to

that no other explanation for mima-mound formation

a spirited exchange of viewpoints.30 Berg acknowledged

seemed applicable to their study area. They also advised

that the seismic hypothesis requires further work to

that this hypothesis should not be applied to all mima-

substantiate, but that it points the way to the solution

like mounds.

of a geologic enigma that has been with us for more

For biologists, the most intriguing hypothesis in

than 100 years. The most recent analysis, by Jennifer

volves burrowing animals, usually pocket gophers. In

Burnham and Donald Johnson in 2012, suggests there

one of the earliest papers on the origin of mima mounds,

could be more than one explanation.31

This page intentionally left blank

Part Four Foothills and

Mountains

This page intentionally left blank

Escarpments and
the Foothill Transition

Chapter 10

The lowlands of Wyoming are interrupted by a dozen

than the adjacent ecosystems. Escarpments appear as

or more mountain ranges (see figs. 1.11.2), all flanked

islands in a sea of sagebrush or grassland, though they

by foothills formed largely from glacial moraines and

can be narrow as well when associated with long ridges

eroding sedimentary rocks. Away from the mountains

(fig. 10.1).1

are numerous escarpments. Both provide distinctive

Foothills and escarpments are important habitats

habitats for plants and animals, adding immeasurably

for big game and other species because of their abrupt

to the aesthetic appeal of the region. The foothills are

topography. Windblown snow accumulates on leeward

comparable to the riparian zone, covering a small, nar-

slopes, which provides a more reliable source of water.

row, elongated portion of the landscape and providing

The lee slopes also serve as windbreaks for big game in

a more moderate environment at some times of the year

the winter, and windward and south-facing slopes are


Fig. 10.1.The coarse, rocky
soil of Cretaceous sandstone
escarpments in the Powder
River Basin supports little
bluestem, ponderosa pine,
Rocky Mountain juniper,
skunkbush sumac, and other
foothill plants. Wyoming big
sagebrush, blue grama, and
western wheatgrass dominate
the shrubland on deeper,
fine-textured soils in the foreground. Elevation 5,000 feet.

155

156 Foothills and Mountains

usually snow-free, providing easier access to forage. In

andin one or two localities in extreme south-central

such places, shelter and forage are available nearby

Wyomingpinyon pine or Gambel oak. Windswept

unlike the mountain forests, where there is abundant

ridges with shallow soils and little snow accumulation

shelter but little winter forage (owing to continuous

have bluebunch wheatgrass and a variety of cushion

snow cover), or the surrounding lowlands, where both

plants (see fig. 8.9).


In general, the foothill-escarpment mosaic is patchy,

shelter and forage may be hard to find.


The vegetation mosaic of escarpments and foothills

a reflection of abrupt changes in rock type, soil depth,

consists of shrublands, grasslands, woodlands, and

snow accumulation, microclimate, slope, and the in

windswept ridges (fig. 10.2). Shrublands dominated

verse texture effect. Vegetation banding is sometimes

by mountain-mahogany are abundant on many rock

observed, such as in the Centennial Valley, west of Lara-

outcrops with little or no soil. On deeper soils where

mie, where different kinds of vegetation are associated

snow accumulates, the common plants are antelope

with bands of sandstone, limestone, shale, siltstone,

bitterbrush, mountain big sagebrush, serviceberry,

and claystone. 2 Similarly, elongated aspen groves can

skunkbush sumac, and snowberry (table 10.1). Black

be found at the contact between impervious granite in

sagebrush or threetip sagebrush commonly occur on

the mountains and the overlying sedimentary strata of

windswept ridges or plateaus with shallow soils, and

the foothills, where runoff from the mountain encoun-

tufted hairgrass and other species sometimes are

ters deeper soils (fig. 10.3). Trees also grow on ridges,

found in leeward arroyos where unusually deep snow-

sometimes because the groundcover there is usually in

drifts develop, sometimes called nivation hollows (see

sufficient to fuel a fire. Thus, trees have a greater chance

fig. 7.8). Foothill grasslands support a sparse cover of

of becoming established than in the adjacent grassland

bluebunch wheatgrass along with other grasses and

or sagebrush steppe, where the competition for water

forbs. The woodlands have varying densities of juni-

from grasses and forbs is more intense and where the

per, limber pine, ponderosa pine, aspen, Douglas-fir,

time between fires is shorter.3

2800

Limber
pine
woodland

ELEVATION (feet)

8000

2600
2400

Juniper
woodland

7000
Basin or
Foothill
grassland

6000

5000

Lodgepole pine,
Douglas-fir, or
Ponderosa pine
forest

Saltbush
desert
shrubland

Sagebrush
desert
shrubland

Sagebrush
steppe

Mixedgrass
prairie

4000
Low

2200

Mixedfoothill
shrubland
or
Ponderosa
pine
savanna
Greasewood
shrubland
Woody
Riparian
draw
woodland

Moderate

High

2000
1800

ELEVATION (m)

9000

1600
1400
1200

WATER AVAILABILITY
Fig. 10.2. Distribution of various kinds of foothill grasslands,
shrublands, and woodlands in relation to gradients of water
availability and elevation in the Bighorn Basin, near the

Bighorn Mountains (Bighorn Canyon National Recreation


Area). Vegetation types that occur above and below the foothills are shown for comparison.

Escarpments and the Foothill Transition

Fig. 10.3. Aspen groves on the east slopes of Centennial


Ridge, in the Medicine Bow Mountains, occur in most ravines
and along the contour where the impermeable granite meets
the permeable soils that have developed on sedimentary
rocks. Sagebrush steppe, dominated by threetip sagebrush,

Plant Adaptations and Vegetation Dynamics

occurs on windswept ridges between the ravines. Douglas-fir


and limber pine occur just above the aspen. Forests at higher
elevations are dominated by lodgepole pine, Engelmann
spruce, and subalpine fir.

quently browsed by deer, especially in the winter. Notably, the shrubs respond by growing lateral branches.

Mountain-Mahogany Shrublands

The more heavily the shrubs are browsed, the more

One of the more conspicuous foothill shrublands is


dominated by mountain-mahogany (Cercocarpus spp.).
This shrub forms dense thickets on rocky or shallow
soils from low elevations on the western Great Plains
(~4,500 feet) up to elevations of about 8,000 feet in
the mountains (fig. 10.4; see table 10.1). Two species
of Cercocarpus occur in Wyoming: the deciduous true
mountain-mahogany in the southern Black Hills and
across the southern half of the state, and the evergreen curlleaf mountain-mahogany in the foothills
of the Bighorn Mountains and also to the west and
south (fig. 10.5). There is little overlap in the distribution of the two species. Both grow on a variety of
bedrock types, including sandstones, limestones, and

spiny the branches become, making further browsing


more difficult. In this way the plants develop their own
defense mechanism.
Some ecologists have observed expansion of curlleaf
mountain-mahogany into adjacent plant communities
during long periods without fire, suggesting that frequent fires historically restricted mountain-mahogany
to rocky sites.6 Curlleaf mountain-mahogany can live
150 years. The shrubs spread by layering or seedling
establishment and can sprout after fires.7 Large seed
crops are common, though often not in consecutive
years.

Juniper Woodlands

shales.4 The abundance of both species is undoubtedly

A characteristic feature of some foothills and escarp-

facilitated by their ability to fix nitrogen; and perhaps

ments are picturesque woodlands dominated by long-

because of their more nitrogen-rich foliage, they are fre-

lived Utah juniper or Rocky Mountain juniper (figs. 10.6

157

Table 10.1.Characteristic plants found in the different kinds of foothill vegetationa

Common name

Latin name

MountainMixed
mahogany
Juniper
foothill
Aspen
Conifer
Oak
shrubland woodland shrubland woodland woodland woodland

Woody
draws

TREES
Aspen

Populus tremuloides

Boxelder

Acer negundo

Bur oak

Quercus macrocarpa

Quercus gambelii

Gambel oakc
Green ash

Fraxinus pennsylvanica

Douglas-fir

Pseudotsuga menziesii

Limber pine

Pinus flexilis

Ponderosa pine

Pinus ponderosa

SHRUBS
American plum

Prunus americana

Antelope bitterbrush

Purshia tridentata

Black sagebrush

Artemisia nova

Mountain big
sagebrush

Artemisia tridentata
ssp. vaseyana

Silver sagebrush
(plains)

Artemisia cana ssp.


cana

Wyoming big
sagebrush

Artemisia tridentata
ssp. wyomingensis

Chokecherry

Prunus virginiana

Common juniper

Juniperus communis

Rocky Mountain
juniper

Juniperus scopulorum

Utah juniper

Juniperus osteosperma

Curlleaf
mountain-mahogany

Cercocarpus ledifolius

True (alderleaf )
mountain-mahogany

Cercocarpus montanus

Snowbrush
ceanothus

Ceanothus velutinus

Douglas hawthorn

Crataegus douglasii

Douglas rabbitbrush

Chrysothamnus
viscidiflorus

Rubber rabbitbrush

Ericameria nauseosa

Saskatoon
serviceberry

Amelanchier alnifolia

Utah serviceberry

Amelanchier utahensis

Skunkbush sumac

Rhus trilobata

Utah snowberry

Symphoricarpos oreophilus var. utahensis

MountainMixed
mahogany
Juniper
foothill
Aspen
Conifer
Oak
shrubland woodland shrubland woodland woodland woodland

Woody
draws

Common name

Latin name

Western snowberry

Symphoricarpos
occidentalis

Wax currant

Ribes cereum

Woods rose

Rosa woodsii

GRASSES
Basin wildrye

Leymus cinereus

Blue grama

Bouteloua gracilis

Bluebunch
wheatgrass

Pseudoroegneria spicata

Idaho fescue

Festuca idahoensis

Spike fescue

Leucopoa kingii

Little bluestem

Schizachyrium
scoparium

Oatgrass

Danthonia spp.

Prairie junegrass

Koeleria macrantha

Sideoats grama

Bouteloua curtipendula

Arrowleaf
balsamroot

Balsamorrhiza
sagittata

Creeping barberry
(Oregon-grape)

Mahonia repens

Field (mouse-ear)
chickweed

Cerastium arvense

Hairy false
goldenaster

Heterotheca villosa

Hookers sandwort

Arenaria hookeri

Lupine

Lupinus spp.

Prairie (fringed)
sagewort

Artemisia frigida

Pricklypear cactus

Opuntia polyacantha

Purple locoweed

Oxytropis lambertii

Pussytoes

Antennaria spp.

Sulfur-flower
buckwheat

Eriogonum umbellatum

Western yarrow

Achillea millefolium
var. occidentalis

FORBS

A dash indicates that the plant is absent or uncommon.

b
c

Bur oak is found only in the Black Hills.

Gambel oak is found only in the Sierra Madre.

160 Foothills and Mountains

Fig. 10.4. (above) Curlleaf mountain-mahogany and a few


ponderosa pine grow on shallow, rocky soils in this area on
the east side of the Bighorn Mountains. Mixed-grass prairie
occurs on deeper soils. True mountain-mahogany is found in
similar habitats in the southern and eastern parts of Wyoming (see chapter 17).
Fig. 10.5. (right) True mountain-mahogany is easily distinguished from curlleaf mountain-mahogany. Drawing by Judy
Knight.

and 10.7). Depending on soil characteristics, associated


plants can include black sagebrush, big sagebrush, antelope bitterbrush, mountain-mahogany, limber pine,
and ponderosa pine (see table 10.1). Rocky Mountain
juniper is common in eastern Wyoming and the north-

tions. However, unlike mountain-mahogany, juniper is

ern Great Plains, where summer precipitation is higher

more capable of expanding into adjacent rangelands on

than farther west. Usually it occurs in ravines or on

deeper soils. Differences in the distribution of juniper

steep slopes, along with yucca, skunkbush sumac, and

and mountain-mahogany may be the result of differ-

ponderosa pine. Utah juniper grows on escarpments in

ences in microclimate, relative ability to spread into

the more arid basins of western Wyoming, for example,

adjacent habitats, or soil-bedrock features that have not

in the Wind River Canyon and Bighorn Basin.

yet been studied in sufficient detail.9 Juniper distribu-

The cause of juniper presence or absence is a puz-

tion in the Bighorn Basin ranges from about 3,600 to

zle, because both species are absent from many habi-

6,000 feet, overlapping with but generally lower than

tats that seem suitable. On the upper or northerly edge

mountain-mahogany shrublands. The lower limits in

of its range, such as in Wyoming, the junipers seem

specific areas seem to be where shallow rocky soils meet

to survive only in warmer thermal belts, where frosts

deeper alluvial or colluvial soils. The upper limit is prob-

are less frequent at night.8 Like mountain-mahogany,

ably controlled by a still-undetermined climatic factor.10

both species are found on a variety of geologic forma-

Most shrublands at low elevations in the Bighorn Basin

Escarpments and the Foothill Transition

Fig. 10.6. (above) Juniper woodland occurs on escarpments


with coarse soils, such as in this area near Bates Creek in the
foothills of the Laramie Mountains, south of Casper. Rocky
Mountain juniper is common in the eastern half of Wyoming;
Utah juniper is common in the western half. Both species are
found near Lander.
Fig. 10.7. (right) The small, scale-like leaves of Rocky Mountain juniper are longer and narrower than the leaves of Utah
juniper. Drawing by Judy Knight.

are dominated by Wyoming big sagebrush, black sagebrush, and broom snakeweed, but curlleaf mountainmahogany commonly is found in the foothills of the
Bighorn Mountains, sometimes with juniper.
Juniper expansion into adjacent grasslands or shrublands, where juniper was previously absent, is one of

that have formed by juniper expansion into former

the most often-discussed ecological topics in the inter-

grasslands or shrublands during the twentieth century

mountain West. A related phenomenon is the estab-

tend to be found on relatively deep soils in regions hav-

lishment of new junipers in formerly sparse juniper

ing summer-dominated precipitation patterns. In con-

woodlandsknown as infilling. Expansion and infill-

trast, the older, persistent woodlands are most common

ing of juniper and pinyon during the twentieth cen-

on shallow, rocky soils and in regions where precipita-

tury have been documented in many areas throughout

tion comes mostly in the winter.12

11

the West. However, the process may be driven by more

What are the ecological processes that drive juni-

than one ecological mechanism, and many woodlands

per expansion and infill? Two mechanisms that have

are stable communities that have persisted with little

received much attention are the effects of livestock

change for hundreds or thousands of years. Woodlands

grazing and fire exclusion. Using tree-ring analysis,

161

162 Foothills and Mountains

rangeland ecologist W. J. Waugh found that juniper

establishment in many parts of the West.19 Increasing

in the Bighorn Basin had been restricted to limestone

atmospheric carbon dioxide may also be facilitating tree

outcrops until about 10 years after the introduction of

growth by increasing water-use efficiency.20 For what-

large herds of cattlea time that also marks the initia-

ever reason, woodland expansion and infill are occur-

tion of fire exclusion. Juniper has since expanded onto

ring in many parts of the West, though the causes may

adjacent sites. Waugh noted that more than 90 percent

not be the same everywhere.

of the juniper seedlings were becoming established

Regardless of the mechanism, herbaceous plant

under sagebrush, which apparently was functioning as

growth is greatly reduced with the establishment of

a nurse plant, and he hypothesized that an increase in

juniper in a grassland or shrubland, owing to more

sagebrush cover following livestock grazing had created

competition for water, light, and nutrients. For a time,

a more favorable environment for juniper invasion.

fires may become less frequent if the shrubs are widely

13

Another important indirect effect of livestock graz-

spaced, thereby perpetuating conditions favorable for

ing is to reduce the fine fuels available for fire. Fires

juniper infilling. However, eventually fuel continuity

at 30- to 60-year intervals formerly prevented juniper

increases and fires become more probable. Prescribed

expansion in parts of the Great Basin of Utah, and

fires are often used to reduce juniper density, but man-

observers have noted that burning can reduce the abun-

agers know that an undesired consequence could be

dance of sagebrush and juniper.14 Neither juniper nor

the invasion of introduced plants, such as cheatgrass

big sagebrush reproduce by sprouting, and woodlands

(see chapter 7). Another concern is that burning could

with old junipers are often restricted to rocky ridges,

reduce the amount of food and thermal cover avail-

where fires are less frequent. Heavy livestock grazing

able for deer and other wildlife. Animals stay warmer at

could also reduce the rate of evapotranspiration from

night under the juniper canopy because of less exposure

the rangeland ecosystem, which would reduce competi-

to the cold sky. Daytime shade during the summer also

tion for water from other plants. Moreover, in combi-

is important.

nation with an unusually wet period, livestock grazing

In northwestern Colorado, on the Colorado Plateau,

could enable the establishment of a large number of

extensive woodlands are dominated by pinyon pine

juniper seedlings. However, this scenario may not apply

and Utah juniper. 21 The two species coexist, but pin-

everywhere, as other observers have observed that

yon pine commonly is found at slightly higher eleva-

drought, in combination with grazing and fire exclu-

tions than Utah junipersuggesting that the juniper

sion, favors juniper invasion.15

is more drought tolerant than the pine. Pinyon pine is

Although fire exclusion and grazing are important

found in only one location in Wyoming, near Flaming

in some places, some young juniper stands represent

Gorge Reservoir in the foothills of the Uinta Mountains.

natural recovery by seedling establishment following

Rocky Mountain juniper is common, often occurring

past fires or other disturbances of mature woodlands.16

with limber pine and ponderosa pine in the foothills. 22

Near the edges of the current ranges of both species,


pinyon pine and the two junipers may still be migrating
into suitable habitats from their Pleistocene refugia. For

Ponderosa PineLimber PineDouglas-fir Woodlands

example, some juniper stands in Wyoming may repre-

Ponderosa pine, limber pine, and Douglas-fir occur com-

sent the first generation to colonize the sites since the

monly in the foothills and on escarpments (fig. 10.8).

last glacial advance.

Infill is also occurring in many

Growing in association with grasslands, juniper, and

stands that contain no evidence of past fire, that is, no

various kinds of shrubland, some of the tree-dominated

basal fire scars on living trees and no charred wood.18

woodlands have been increasing in density and expand-

Some of these woodlands have also been grazed only

ing their rangepossibly as a result of fire exclusion,

lightly or not at all. In such places, fire exclusion and

although climate conditions in the twentieth century

grazing cannot be the drivers of infill. Climate variation

have been favorable for these tree species just as for juni-

probably is more important. Several warm, moist peri-

per and pinyon.23 Increasing densities of ponderosa pine

ods during the 1900s were especially favorable for tree

can pose problems for ranchers, primarily because the

17

Escarpments and the Foothill Transition

Fig. 10.8.Escarpments and foothills on the eastern plains


typically have ponderosa pine woodlands, such as along the
Hat Creek Breaks, north of Lusk (Oligocene and Miocene
sandstones and claystones). This ridge is part of Pine Ridge,

which extends westward to north of Casper and eastward into


southern South Dakota. Mixed-grass prairie is also common,
sometimes with silver sagebrush and Wyoming big sagebrush.
Elevation 5,000 feet.

additional trees can reduce the amount of forage and

Although ponderosa pine is often common on escarp-

increase evapotranspiration (thereby reducing stream-

ments in eastern Wyoming, where deep water percola-

flow), and also because the pine needles can be toxic if

tion is presumably possible, some geologic substrata

their livestock are hungry enough to eat them.

provide better conditions for tree growth than others.

24

In Wyoming, ponderosa pine is found primarily in

Wyoming ecologist Jim States found that the growth

comparatively warm areas with higher summer pre-

of ponderosa pine was best on Mowry shale, where the

cipitation. Such environments are found in the eastern

trees had wider and less variable annual ring widths in

half of the state, primarily in the Black Hills and Bear

their wood.27 Low annual variability in ring width sug-

Lodge Mountains, at lower elevations on the east and

gests that this marine shale, despite the variable climatic

north slopes of the Bighorn Mountains and the Lara-

conditions from year to year, provided a consistent envi-

mie Mountains, in a few localities around the Medi-

ronment for growth. In contrast, trees on the Pine Ridge

cine Bow and Seminoe mountains, and on lowland

and Cloverly sandstones had narrower and more variable

escarpments scattered throughout the Great Plains. 25

annual rings, even though they grew near the Mowry

Associated plants include skunkbush sumac, Rocky

shale and surely had the same climatic conditions.28 The

Mountain juniper, and two grassessideoats grama

association of some foothill species with fracture lines in

and little bluestem (see table 10.1). Ponderosa pine is

the bedrock can be striking (see fig. 3.3).

essentially absent from western Wyoming, perhaps

Limber pine sometimes occurs with ponderosa

because of a growing season that is currently too short

pine, but it is more typical where the climate is drier

and dry. It does occur in the foothills of the Uinta

and colder. Unlike ponderosa, it extends upward to the

Mountains of Utah (such as in Dinosaur National

alpine treeline. The species tends to be restricted to rocky

Park), where the elevations are lower and the Arizona

soils and ridges, where the seedlings have relatively

Monsoon provides slightly more summer precipitation

little competition from other plants.29 It is commonly

than in Wyoming. 26

found on the lee side of boulders, where the seedlings of

163

164 Foothills and Mountains

tains just west of Yellowstone National Park, and in


other parts of the Greater Yellowstone Ecosystem (see
chapter 15). Compared to the eastern foothills, a larger
proportion of the annual precipitation occurs in the
winter, which apparently favors Douglas-fir. The soils
typically are derived from sedimentary rocks and are
more fertile. 30 Limber pine is sometimes associated with
Douglas-fir in western Wyoming, such as in Fossil Butte
National Monument.
The dynamics of Douglas-fir woodlands is much the
same as for ponderosa pine woodlands growing in similar situations, commonly with expansion into adjacent
grasslands and shrublands during long fire-free intervals.31 In parts of Montana, scars on trees have revealed
that fires occurred every few decades before 1890, confining the trees to rocky sites or the lee sides of slopes,
where the fires burned less frequently. Comparable
results were found for ponderosa pine woodlands in
Devils Tower National Monument, where, from 1770 to
1900, the average number of years between fires was 14.
Since 1900, fires have been much less frequent, about
once every 40 years, owing to fire suppression and the
Fig. 10.9. Limber pine occurs on rocky escarpments from low
elevations near Pine Bluffs in the southeast up to the alpine
treeline of some mountains in eastern Wyoming. In northwestern Wyoming, whitebark pine is found in the alpine
zone. This photo, taken in the Laramie Range, shows widely
scattered limper pine in a high-elevation shrubland dominated
by threetip sagebrush, fringed sagewort, junegrass, mountain
muhly, and Sandberg bluegrass. Elevation 8,000 feet.

great reduction in fires ignited by humans.32

Mixed Foothill Shrublands


Sagebrush steppe extends into the foothills in many
areas, often intermingled with mountain-mahogany
shrublands and woodlands dominated by pines,
Douglas-fir, and juniper. Species composition changes
as the environment becomes cooler and relatively more

the now-mature trees would have been protected from

moist. Mountain big sagebrush becomes the domi-

excessive wind damage. Once established, however,

nant variety instead of Wyoming big sagebrush, and

limber pine occurs on some of the windiest sites in the

other shrubs are common, namely, antelope bitter-

region (fig. 10.9). Indeed, the unusually limber branches

brush, common chokecherry, serviceberry, skunkbush

that give this tree its name probably are an adaptation

sumac, snowberry, snowbrush ceanothus, wax currant,

to minimize wind damage. Associated species include

and Woods rose (fig. 10.11; see table 10.1). Great Basin

various shrubs, such as common chokecherry, ground

wildrye and antelope bitterbrush grow in ravines and

juniper, mountain big sagebrush, Oregon-grape, wax

on sites where snow accumulates. Common grasses

currant, and western snowberry, along with forbs and

and forbs include bluebunch wheatgrass, Idaho fescue,

grasses (for example, spike fescue).

prairie junegrass, arrowleaf balsamroot, hairy golden

Foothill woodlands in the western half of Wyoming

aster, and lupine.

are dominated by Douglas-fir rather than ponderosa

Serviceberry is a conspicuous shrub in some mixed

pine (fig. 10.10). Many of the best examples are found

foothill shrublands. Two species are commonly rec-

in the Sunlight Basin, the Lamar Valley of Yellowstone,

ognized in Wyoming: Saskatoon serviceberry, which

Jackson Hole, the lower slopes of the Centennial Moun-

occurs throughout the state, and Utah serviceberry,

Escarpments and the Foothill Transition

Fig. 10.10. (above) Douglas-fir is the most common tree on


north slopes at lower elevations in the Greater Yellowstone
Ecosystem. The south slopes are generally too dry for trees
but support a mixed foothill shrubland with bluebunch
wheatgrass, big sagebrush, Saskatoon serviceberry, and other
plants. Mountain big sagebrush dominates the shrublands
in the foreground. Photo taken looking west toward Hoback
Junction.
Fig. 10.11. (left) Saskatoon serviceberry, antelope bitterbrush,
rabbitbrush, and mountain big sagebrush dominate this mixed
foothill shrubland in the Sierra Madre. Aspen grows in the
ravine. Elevation 8,100 feet. Photo by Ken Driese.

which is found to the west and south.33 Both species


are deciduous, capable of root sprouting, and are important browse plants for deer and elk. At lower elevations,
where the frost-free period is longer, skunkbush sumac
is common, especially on the east side of the Bighorn
and Laramie mountains.
The relatively moist environment of the foothills allows for greater plant cover, which enhances
the probability of fires, whether caused by lightning
or humans. Periodic fires in some areas prevent the
growth of juniper and other conifers. Temporary shifts

165

166 Foothills and Mountains

in species composition may also occur following fire.

with succession or if grazing is confined to the period

Most of the shrub species have the ability to resprout

when the plant is dormant.37

from surviving roots and other belowground structures

Some foothill grasslands occur on extraordinarily

(big sagebrush and juniper are notable exceptions).

windswept plateaus or slopes with shallow soils. The

Fires commonly expose mineral soil, facilitating the

grasses are scattered and small forbs are common.

re-establishment or addition of some species through

Many of the forbs have the cushion plant growth form

seedling recruitment. Antelope bitter


brush, snow-

(see fig. 8.9), which keeps the stems and leaves densely

brush ceanothus, skunkbush sumac, and mountain-

aggregated at the soil surface, where the air is warmer.

mahogany are capable of nitrogen fixation because

Cushion plants are typical of windswept areas in alpine

of the actinorhizae formed by bacteria in the genus

tundra as well as in the lowlands, though the species

Frankia. This adaptation enables the establishment of

are different.38

these species on relatively infertile soils. Snowbrush


ceanothus is capable of layering, but in addition, seed
dormancy in this shrub is broken by the heat of a fire;
seedlings commonly are abundant in burns where the
species existed previously.
Greasewood occurs in a few places in the foothills,

Deciduous Woodlands
Four other vegetation types in the foothill-escarpment
mosaic are woodlands characterized by different combinations of aspen, chokecherry, Gambel oak, and bur

an interesting anomaly, as it is usually found where


groundwater is near the surface on saline soils in desert shrublands. The presence of greasewood on escarpments has been attributed to porous lignite seams that
cause groundwater to move to the surface on hillsides.34
Seeps or perched water tables, with concomitant salt
accumulation as some of the water evaporates, typically
develop on such sites.

Foothill Grasslands
Foothill grasslands can be found on windy slopes or plateaus, where snow does not accumulate in large quantities, soils are too shallow for most shrubs, or summer
rainfall is higher. Bluebunch wheatgrass is the characteristic plant, occurring most often on relatively warm,
dry sites. Idaho fescue is typical of the higher, more
mesic montane grasslands (above 6,900 feet).35 Threetip
sagebrush, a dwarf shrub, occurs with Idaho fescue on
windswept, shallow soils (see fig. 10.9). Little bluestem
is sometimes found on the slopes of escarpments at low
elevations.36
Bluebunch wheatgrass is a highly preferred forage
species for livestock and big game. At lower, warmer
elevations, it is usually found on north slopes or near
snowbanks, if it occurs at all. Various investigators have
observed that bluebunch wheatgrass becomes less common after hot fires or heavy grazing in spring or summer. Cheatgrass is a common invader following such
disturbances, but bluebunch wheatgrass increases again

Fig. 10.12. Ravines in the foothills of the Bear Lodge Mountains and Black Hills near Sundance typically have chokecherry, skunkbush sumac, Woods rose, and other shrubs.
Shrublands such as this are sometimes referred to as woody
draws. Mixed-grass prairie occurs on the upland, with ponderosa pine on some ridgetops. Elevation 3,900 feet.

Escarpments and the Foothill Transition

Fig. 10.13. Aspen groves occur in some places on the lee


side of the Continental Divide south of Rawlins, where deep
snowdrifts develop during winter. The snow provides ade
quate water for shrubs and trees in an area that elsewhere is
semi-arid. This photo shows a portion of a doughnut-shaped
grove, known as an aspen atoll, where the center is dominated by shrubby cinquefoil, mountain silver sagebrush,

lupine, tufted hairgrass, and various other grasses and forbs.


Deep snow that persists into early summer prevents tree
establ ishment in the center. The shrubs in the foreground
and on the left are old aspen sprouts that are regularly
broken by the heavy snowpack. Aspen trees develop just
beyond the shrubby aspen, where snow accumulation is less.
Elevation 7,600 feet.

oakall of which are deciduous and occur on mesic

Aspen-dominated woodlands tend to occur where

sites with deeper soils. The additional moisture appar-

the soils are deeper, such as on lower slopes where water

ently comes from a higher soil water-holding capacity,

seeps to the surface or on the lee sides of ridges where

more snow accumulation (from drifting), more summer

snow accumulates. Associated plants include species

precipitation, or runoff from the slopes above.

that occur in aspen woodlands at higher elevations (see

Chokecherry woodlands, also known as hardwood

chapter 11). Sometimes the aspen completely surround

draws or woody draws, are commonly part of a veg-

a snowdrift, forming a doughnut-shaped grove around

etation mosaic on the eastern plains that includes

a meadow that has been referred to as an aspen atoll

grassland and woodlands dominated by ponderosa

(fig.10.13). Such groves develop when trees first become

pine (fig. 10.12). In addition to chokecherry, common

established near ridgetops, which causes further accu-

shrubs include American plum, skunkbush sumac,

mulation of snow downwind. If the snow is deep

silver sagebrush, and western snowberry. 39 Trees com-

enough, it persists until midsummer and only herba-

monly found with the shrubs are boxelder and green

ceous plants capable of surviving with a short growing

ash (see table 10.1). Except for the lack of cottonwood,

season can persist. Gradually the aspen grow completely

chokecherry woodlands have the characteristics of

around the snowdrift and meadow by root sprouting.

some riparian woodlands. Specifically, they have more

The trees also expand inward, toward the edge of the

plant biomass than is found on adjacent shrublands,

central meadow. Here the average snow accumulation is

which provides good wildlife cover as well as forage

sufficiently shallow to allow the aspen to persist, but the

and shade for livestock.

weight of the snowpack causes breaking and bending,

167

168 Foothills and Mountains

resulting in trees that look more like shrubs. Only on


the outer fringes of the clone do aspen sprouts grow tall.
Bur oak woodlands are found in the northern and
eastern Black Hills (see fig. 16.6), and Gambel oak
woodlands can be found on the west side of the Sierra
Madre, on the Wyoming-Colorado border. Late spring
frosts and summer drought most likely inhibit Gambel
oak farther to the north by preventing establishment
of oak seedlings.40 Southward in Colorado and Utah
(see chapter 3), where Gambel oak becomes increasingly common, late-spring frosts are less frequent, and
summer precipitation is higher because of the Arizona
Monsoon. Gambel oak occurs in northern Utah primarily where temperature inversions regularly create
warmer thermal belts in the foothills, with an average
of 90 frost-free days a year.41 Here, Gambel oak is usually found on sloping upland sites between about 5,500

Plant-Animal Interactions in the Foothills


Escarpments and foothills provide habitat for a wide
diversity of wildlife. Of particular interest are elk and
other big game species, which tend to winter in these
landscapes, where snow cover is less and food more
available than higher in the mountains. Many winter
ranges have been greatly altered by agriculture, exurban
development, and other human activities. This loss of
habitat, coupled with the fact that winter ranges are less
extensive than the associated summer ranges, means
that animals are concentrated in small areas for long
periods each winter. Management of ungulate winter
ranges has long been controversial in Wyoming and
much of the intermountain West

Aspen, Elk, and Fire

and 7,500 feet, above the pinyon-juniper zone but

The management of foothill aspen groves has been

below ponderosa pine and aspen. Curiously, bur oak

one of the most interesting and highly debated natural

occurs only in the northeastern corner of the state

resource issues in the region.44 The controversy began

farther north, but the elevation is lower (and warmer)

in the mid-1950s, when some aspen groves in north

and annual precipitation is comparatively high (see

western Wyoming appeared to be dying (fig. 10.14). Such

fig.3.4). If it were not for late-spring frosts, oak would

groves were thought of as decadent. Aspen is unique,

probably be a co-dominant with ponderosa pine else-

because the numerous trees in a grove result from the

where in Wyoming, as is common in Colorado (for

root sprouting of only one or a few plants. Each shoot

example, near Colorado Springs). Oak might also occur

aboveground is better thought of as a branch. Seedling

with aspen and chokecherry. Climate change could

establishment is rare in the foothill environment, but

lead to an ex
pansion of oak woodlands if relatively

when successful, the roots sprout additional stems that

warm springs coincide with sufficient summer precipi-

grow into what most people think of as new treesa

tation for the seedlings to survive.42 Once established,

classical case of cloning that is more common in her-

oak tolerates cold, dry conditions.

baceous plants. Like other species of Populus, each stem

Oak woodlands burn readily during dry seasons, but

is comparatively short-lived, with few surviving past

the roots are not usually killed and they sprout vigor-

150 years. However, the root system lives indefinitely,

ously, often leading to an increase in density.43 Other

producing new stems around the edges of groves or

species invade during long fire-free intervals, such

where older stems are damaged.

as Rocky Mountain juniper and ponderosa pine (and

The apparent loss of foothill aspen groves is attrib-

pinyon pine, white fir, and canyon maple in north-

uted to two factors: fire suppression and the excessive

ern Utah). Both Gambel and bur oak woodlands pro-

browsing of sprouts by large mammals, particularly

vide grasses and forbs for big game during the winter,

elk.45 Browsing was the first factor to be debated.

and deer, elk, wild turkey, and squirrels consume the

Many of the so-called decadent groves in northwest-

energy-rich acorns. With time, forage production under

ern Wyoming occurred along known elk migration

oak canopies declines because of shading and competi-

routes, and many were located adjacent to winter

tion for water. Fire can be used to stimulate the growth

feeding groundsnotably the National Elk Refuge in

of more forage, but ranch buildings, summer homes,

Jackson Hole. Though elk are grazers in the summer,

and resorts located in the foothill environment can pre-

they browse on aspen sprouts in winter and spring.

vent this otherwise sound practice.

Too much browsing for too many consecutive years

Escarpments and the Foothill Transition

reduces the number of sprouts. When aspen groves are


fenced to exclude large herbivores, the small sprouts
often grow into trees (fig. 10.15).
An interesting perspective on decadent aspen stands
was presented by U.S. Forest Service ecologist Norbert
DeByle.46 He suggested that one of the causes is political
pressure to maintain elk populations at the same high
level year after year, to satisfy hunters, guides, and game
and fish agencies that depend on license sales. DeByle
reasoned that in the 1800s the number of elk fluctuated considerably, largely stemming from higher mortality during some winters than others, and that aspen
sprouts had a better chance of developing into trees
during periods when the browsing pressure was low

Fig. 10.14. (right) An aspen grove in the Greater Yellowstone


Area that appears to be dying, most likely because of excessive
browsing by elk (see chapter 15). Mountain big sagebrush,
arrowleaf balsamroot, and other foothill plants occur in the
area; lodgepole pine and limber pine can be seen in the background. Elevation 7,100 feet.
Fig. 10.15. (below) Aspen groves usually recover quickly when
protected by a fence from browsing, such as in this area in
northwestern Wyoming. Aspen exists outside the fenced
exclosure, to the right, but only as small, heavily browsed root
sprouts. See chapter 15.

169

170 Foothills and Mountains

(anatural rest rotation). Elk populations still fluctuate,

and drier, as discussed in chapter 11. The magnitude

but probably to a lesser extent because of winter feed-

of the problem is not well known. What proportion of

ing programs. Notably, the number of elk has declined

the aspen groves in different mountain ranges is in a

during the past 12 years or so, almost certainly because

degraded condition? Could some aspen groves have

of the reintroduction of gray wolves in 1995 combined

been in poor condition long before the arrival of

with droughts. Furthermore, some evidence suggests

EuroAmericans in the 1800s?

that the wolves harass elk herds from time to time, causing them to move rather than stay in one aspen grove
for a long period. Aspen could be one of the beneficia-

Winter Rangelands for Deer and Elk

ries of restoring the regions top carnivore (but see chap-

The winter survival of pronghorn, deer, elk, and moose

ters 12 and 15 for further discussion of this topic and

depends largely on adequate food during the summer

sudden aspen decline).

and fall for the development of fat reserves sufficient to

Wildlife managers are sometimes sensitive about elk

sustain the animals during long, harsh winters. Deep

receiving so much of the blame for aspen groves that

snow or unusually cold winters cause a more rapid

appear to be in bad condition. They argue, correctly in

depletion of fat reserves, but fortunately, the snow com-

some cases, that livestock frequently congregate under

monly melts rather quickly on south slopes in the foot-

the trees because of the abundant forage that develops

hills. Also, the rugged topography makes it relatively

in the relatively moist environment where aspen is

easy for the animals to find windbreaks. The bark and

typically found. Cattle and sheep can cause consider-

twigs of trees and shrubs, though generally not pre-

able damage to aspen through browsing, trampling, and

ferred, often are consumed during this time, partly

bedding.

because they are visible above the snow. Browse lines

Fire is another factor that cannot be ignored in ex

develop on junipers, willows, and other tall shrubs but

plaining the apparent decline in some aspen groves.

are not necessarily an indication of excess numbers of

University of Colorado ecologists Thomas Veblen and

wintering animals at the present time (fig. 10.16).

Diane Lorenz, working in Colorado, suggested that

By the time spring arrives, winter ranges may seem

widespread burning in the late 1800s and early 1900s,

to be heavily grazed and browsed, yet the plants typi-

whether caused by lightning or humans, might have

cally recover quickly, because the herbivory occurs

triggered the initiation of new stands of aspen over

largely when they are dormant. Depending on manage-

large areas.47 Those stands are now aging, which could

ment practices and the extent of overlap between the

explain why the groves are failing over large areas. How-

summer plant preferences of livestock and the winter

ever, aging stands could also be the result of fire sup-

preferences of big game, winter range plants may have

pression during the past 60 or 70 years. Possibly more

a full summer of rest and become senescent before they

important, with fire suppression conifers have become

are again subjected to large ungulates. The seasonal

dominant in some groves, so that the fewer remain-

movements of the animals create a natural rotation sys-

ing aspen are subjected to more browsing than before.

tem.48 More than one year of rest may result if the herd

Large-scale fires could stimulate the development of

selects a different part of the foothills for wintering or if

new aspen forests, providing a larger food base and

high mortality during a particularly severe winter leads

removing some of the browsing pressure on any partic-

to fewer animals during the following winter.

ular grove. Notably, many new aspen seedlings became

A primary concern of wildlife managers during the

established in the Greater Yellowstone Area after the

past 10 years has been the decline in mule deer. Long

1988 fires (see chapter 15).

periods of droughtpossibly climate-change-induced

There is still much debate on the relative significance

droughtprobably have been a contributing factor,

of elk browsing, fire suppression, and wolf predation on

because inadequate moisture limits the summer forage

the condition of aspen woodlands in the foothills. Cli-

available to build up fat reserves. Another factor has

mate change could be another contributing factor, pos-

been winter habitat fragmentation by exurban devel-

sibly reducing aspen vigor if summers become longer

opments. All such factors reduce the amount of forage

Escarpments and the Foothill Transition

transmission (such as brucellosis or chronic wasting disease), and the fostering of a public perception that protecting suitable winter habitat is not important.49 The
benefits are perceived to be fewer agricultural conflicts,
larger herds for hunters (and increased license sales for
the wildlife agencies), and relief from the angst associated with knowing that many animals are starving
during harsh winters because of industrial activity and
rural subdivisions.

Pine Seed Dispersal by Birds


One of the most interesting plant-animal interactions
in foothill and escarpment landscapes is that between
certain seed-eating birds and limber and pinyon pines.
Unlike most conifers, these pines have comparatively
large nutritious seeds, and the cone scales tend to hold
the seeds in place, even after the cones have opened.
Not surprisingly, the seeds are an important source of
food for various birds, squirrels, and even grizzly bears.
The pines benefit by having their seed dispersed and
sometimes planted by birds and squirrels.
The best-studied example is the Clarks nutcracker,
a common bird in the foothills and mountains (fig.
Fig. 10.16. Browse lines, such as on the taller, older Rocky
Mountain juniper in the background, are formed when
adequate food is not available for elk during the winter. Such
browse lines might have existed for centuries near elk and
deer winter ranges. The two junipers in the foreground are
younger and do not appear to have been heavily browsed,
most likely reflecting reduced populations of the wintering
elk and deer during their development. Photo taken between
Gardiner, Montana, and the border of Yellowstone National
Park. Elevation 6,000 feet.

availability directlyand indirectly as well, because


human activity forces the ungulates into areas where
the amount and quality of winter forage is either marginal or where another herd already resides.
Recognizing the problems associated with reduced
winter range, federal and state agencies initiated winter
feeding programs (see chapter 15). From the beginning,
the wisdom of feeding elk and deer in the winter has
sparked controversy, with Wyoming, Idaho, and Utah
having state-sponsored feed grounds, whereas Montana
and Colorado do not. Aside from rather high costs per
animal, the problems are excess herbivory of trees and
other plants in the vicinity of feeding grounds, disease

10.17).50 In Arizona, a single nutcracker was observed


carrying 95 pinyon seeds in a pouch behind its beak for
a distance of 13 miles. Usually the seeds are collected in
late summer and buried on south-facing slopes, which

Fig. 10.17.Clarks nutcrackers bury the peanut-sized seeds of


limber and whitebark pine in the soil, often on warm south
slopes. Seeds that are not recovered often develop into new
trees, some of which grow in clusters if a bird deposited more
than one seed in a single hole. Photo by LuRay Parker/
Wyoming Game and Fish Department.

171

172 Foothills and Mountains

are often snow-free during late winter. Many seeds are

sunny slopes that are frequently snow-free in winter,

relocated in the spring, a time coinciding with the feed-

conditions that might reflect an instinctive preference

ing of nestlings and when food for seed-eaters is in short

of the bird. Seeds cached on such sites would be more

supply. However, some seeds are left in the ground and

accessible in the winter, when food is scarce. Burned for-

develop into new treescommonly clumps of trees, one

ests are also selected by the birds, which could increase

from each of several seeds that were cached by the bird

the probability of limber and whitebark pine becoming

in the same hole. In a similar manner, the nutcrackers

part of the new forest that develops. Its well known that

also disperse the seeds of limber pine and whitebark

grizzly bears in the Greater Yellowstone Area consume

pine.51

large numbers of whitebark pine seed in the fall (see

As might be predicted, the distribution of nut

chapter 15). From the bears perspective, the more seeds

crackers overlaps with that of pines having large, wing-

planted by birds, the betterespecially at a time when

less seeds (in Europe as well as North America). The

many of the adult trees are dying from the depredations

plants benefit by having their seeds dispersed more

of mountain pine beetles and the invasive, non-native

widely and even planted. In fact, the large wingless

white pine blister rust. Conservation biologists are

seeds cannot be dispersed effectively over long dis-

attempting to maintain populations of whitebark pine

tances by any mechanism other than bird transport.

by planting seeds from trees that appear to be resistant

Seeds that simply fall to the ground are more likely

to the disease. (See chapter 14 for further discussion of

to be eaten by small mammals, and the seedlings are

white pine blister rust.)

less likely to become established on the litter layer of


the forest floor, a less favorable environment than that

Overall, the exposed bedrock of foothill and escarpment

provided by burial in mineral soil. Moreover, seedlings

landscapes creates sharp environmental boundaries and

that originate from fallen seed may be less likely to

a patchy vegetation mosaic. Such landscapes are highly

survive because of competition with the nearby and

valued real estate for people, providing the amenities of

already well-developed parent tree. This coadaptation

a moderate, low-elevation environment close to pictur-

between birds and some pines is further illustrated

esque escarpments that provide windbreaks, together

by the observation that the nutcracker distinguishes

with grand vistas of the plains below and the mountains

between viable and nonviable seeds. There is little to

above. Dispersed residential development has been the

be gained for either the bird or tree if less nutritious,

fastest growing land use in the United States since 1950

nonviable seeds are cached.

and has resulted in habitat fragmentation throughout

Bird dispersal appears to be one of the factors caus-

the intermountain Westoften in the foothills. As dis-

ing the patchy distribution pattern of limber and white-

cussed further in chapter 18, debates about how best to

bark pine. Typically, these trees are found on ridges or

use these lands surely will continue.

Chapter 11

Mountain Forests

Rising above the plains, basins, and foothills are moun-

uneven snow accumulation and melting, with some

tains that greatly diversify the landscapes of the region.

patches of ground becoming snow-free much earlier

Most are anticlinal or fault block in origin and have

in the spring than others. Some species, such as dwarf

cores of Precambrian granite, gneiss, and quartzite. In

huckleberry, cannot tolerate snow cover that lasts until

some areas the Precambrian rocks are still covered by

early July.1

sedimentary strata formed at the bottoms of ancient seas

The most noticeable environmental gradient in

that spread across Wyoming millions of years ago, before

mountains is a decrease in temperature with an increase

mountain building began. More than 50 peaks rise

in elevation (see fig. 3.6). The rate at which tempera-

above 13,000 feet, mostly in the Wind River Mountains.

ture cools can be 25F for every 1,000 feet of elevation

Some of these mountain ranges are massive (see fig. 1.1),

gain, depending on the time of year.2 The variability

whereas others are small islands in a sea of sagebrush.

also depends on differences in mountain extent, cold-

Glaciers have shaped portions of the higher mountains.

air drainage, snow cover duration, potential for tem-

Large or small, the mountain ranges are important


economically. Almost all of the forests are found there,

perature inversions, and other climatic and topographic


factors.

and most of the water used by agriculture, industry,

Other environmental factors also change with ele-

and municipalities originates there as snowfall. The

vation. The snow-free period and length of growing

mountains are also important for outdoor recreation.

season are typically shorter high on the mountains,

Ecologically, they are interesting because of their eco-

but precipitation generally increasesat least to mid-

systems and the ways that they affect the lowlands (for

elevations, where snow accumulation is greatest. Snow

example, through the development of alluvial fans and

accumulation combined with lower evaporation rates

rainshadows). Indeed, the landscape patterns observed

caused by cooler temperatures provides an environment

in the foothills and lowlands are caused by the nearby

that, in general, is considerably wetter than in the sur-

mountains.

rounding lowlands (see fig. 3.4). Countering the effect


of cooler temperature is lower atmospheric pressure,

Variation in Mountain Environments


Spatial variation in mountain forests can be attributed

which allows relatively high rates of evapotranspiration.3 Still, plant water stress is less frequent in midelevation forests than in any other upland ecosystem.

partially to environments that change with elevation,

The effects of topography are just as important as

slope steepness and aspect, and the influences of under-

those of elevation. South-facing slopes are warmer and

lying bedrock on soil development. Other factors are

drier and north-facing slopes are cooler and wetter

173

174 Foothills and Mountains

than might be expected based solely on elevation. Val-

all interact to determine the spatial distribution of any

ley bottoms are often cool because of cold-air drainage.

given plant species.4

Thus, species typically found high on the mountains


may extend to low elevations in ravines or on north
slopes with less direct sunlight, and species requiring a

Surviving in the Mountains

warmer environment may be found at unusually high

Perhaps the primary challenge for plant survival in the

elevations on south slopes (fig. 11.1).

Rocky Mountains is the short, cool, and sometimes

Soils also have a pronounced effect on the landscape

dry growing season. Several distinctive adaptations are

mosaic. For example, meadows occur in the Bighorn

involved. Most notable is that photosynthesis occurs

Mountains on comparatively dry, fine-textured soils

in montane plants at temperatures near freezing, or

even at high elevations where forests might be expected

sometimes even below,5 and many plants are evergreen

(see chapter 12). Forests of Douglas-fir, lodgepole pine,

or wintergreen (that is, they have chlorophyll in their

Engelmann spruce and subalpine fir grow on other

leaves or stems all year). The ability to tolerate low tem-

kinds of soils (Inceptisols and Alfisols), where moisture

peratures, combined with the presence of chlorophyll

stress is less severe in late summer (such as on north

throughout the year, enables some plants to extend

slopes). Lodgepole pine is typically the most common

their growing season into late fall and begin the next

tree in forests on coarse, less fertile soils derived from

one earlier in the spring. Evergreen conifers are excel-

granite or rhyolitic lava. Multiple factorssoils, local

lent examples, but other plants have similar adapta-

climate, influences of herbivores and pollinators, and

tions. For example, such deciduous plants as aspen

local history of disturbance and species migrations

and dwarf huckleberry have chlorophyll in their stems,


which permits photosynthesis even when they have no
leaves. The green stems of the huckleberry may account

Fig. 11.1. Approximate distribution of different forest types in


relation to elevation and moisture availability. Note that each
kind of forest typically occurs at the lowest elevation in cool,
moist ravines. Foothill shrublands occur just below woodlands and forests at lower treeline.

Some shrubs and herbaceous plants have leaves

Wet
meadow

Fellfield

Turf

Willow

enough to permit light penetration, when the temperature is at or near freezing.7 The capacity for photosyn-

Spruce-fir
Lodgepole pine
Aspen

9000

thesis at cold temperatures is an important adaptation,


3000

Aspen

9000

Lodgepole pine
Willow
8000
Alder
Birch

2500
Aspen

Douglas-fir
or
Ponderosa pine

Foothill
shrublands

but simply being able to survive the cold is another. A


comparison of low-temperature tolerances of numerous
tree species found that, unlike the trees from warmer
climates, those of the Rocky Mountains could tolerate
temperatures of 76F.8
Another problem for mountain plants is acquiring
nutrients from infertile, coarse-textured soils. All of the
tree species have mycorrhizae (fig. 11.2), which are con-

7000
Cottonwood

synthesis and growth as soon as the snow is shallow

ELEVATION (m)

ELEVATION (feet)

deciduous evergreen plants.

elk sedge and wintergreen. They are capable of photo-

3500

Alpine

6000

synthesis.6 Both aspen and huckleberry are examples of

that remain green even under the snowfor example,

12000

9000

for a substantial portion of the plants annual photo-

centrated in the top 12 inches of the soil, where limiting


2000

nutrients are most likely to be available. Once incorporated into plant tissues, the nutrients are retained for

Wet
Cool North
valley ravine slope
bottom

E-W
slope

South Ridge
slope

MOISTURE GRADIENT

extended periods simply by the longevity of leaves and


twigs. Lodgepole pine needles, for example, persist for
518 years, depending on environmental conditions.

Mountain Forests

Several studies have shown genetic differentiation over


short environmental gradients in Rocky Mountain
conifers, and aspen groves at high elevations are genetically different from those in the foothills.12 Considerable genetic variability, one aspect of biodiversity, can
be expected in landscapes with rapid spatial and temporal changes.
Vertebrates have an equally diverse range of adaptations for surviving in mountain landscapes. Some
migrate in the fall to warmer or less stressful environments at lower elevations or farther south, minimizing their exposure to cold temperatures, deep snow,
Fig. 11.2. Microphotograph showing swollen mycorrhizal root
tips on lodgepole pine. The filamentous hyphae of the fungus
extend from the bulbous tips into the soil, facilitating the
uptake of water and nutrients. Photo by Steven L. Miller.

and food shortages during the winter. Other species


survive because of remarkable insulation provided by
fur, fat, and feathers, and because they find or build
insulated shelters, such as in deep snow or under logs.
Energy stored during the preceding summer in fatty
tissue or food caches is especially critical for animals

When the leaves of evergreen plants are about to fall,

that are year-round residents. Many species have rel-

their limiting nutrients are first partially reabsorbed by

atively large feet that enable swift travel across deep

the twig.9 Thus, plants in nutrient-deficient environ-

snow, for example, red squirrels, martens, lynx, and of

ments have evolved mechanisms for conserving the

course, snowshoe hares.13 A few are able to enter a state

limiting nutrients they already have.

of torpor, or reduced physiological activity brought

Short, cool growing seasons also present problems for

about by decreased metabolism and body temperature.

seedling establishment, because plants that are only a few

Examples include bears, marmots, bats, reptiles, and

months old are the most susceptible to water, tempera-

amphibians.14

ture, and nutrient stresses. Seeds often germinate despite


cool temperatures in the spring, allowing the maximum
time possible for seedlings to develop the stored energy

The Forest Community

and cold-hardened tissues needed for the following win-

Forests are characterized by trees, but, with regard to

ter. Moreover, the formation of mycorrhizal roots typi-

the numbers of species, most plants in the commu-

cally occurs in the first few weeks after germination.10

nity are shrubs and herbaceous plants in the under-

Despite these adaptations, seedlings often succumb to

story. Common shrubs include buffaloberry, dwarf

summer drought, late-spring frosts, and other factors.11

huckleberry, and ground juniper; the herbs include

Because seedling establishment is so precarious, nearly

heartleaf arnica, lousewort, pyrola, and various grasses

all mountain plants are perennials that live for many

and sedges (table 11.1). Many animals and legions of

years.

microorganisms coexist with the plants. In one way or

Different species have evolved to tolerate differ-

another, all are adapted to living with the other spe-

ent levels of environmental stressa fundamental

cies, large and small. Indeed, some are dependent on

principle in considering the causes of patches in land-

others. A special example consists of plants known as

scapes anywhereand stress levels change rapidly

mycoheterotrophsthose that lack chlorophyll and

with changes in elevation, soil, and topographic posi-

obtain their energy from decaying tree roots by way

tion. Consequently, species composition often changes

of saprophytic fungi. A common example in Rocky

abruptly in the mountains. Variation occurs even with-

Mountain forests is pinedrops (Pterospora andromedea).

out changes in species composition, as many species

Clearly, understanding forest ecology is not possible by

have a wide range of genetically determined tolerances.

studying the trees alone.

175

Table 11.1 Some characteristic plants of mountain forests and woodlands in Wyominga

Common name

Latin name

Limber
Ponderosa
pine
pine
Douglas-fir
woodland
forest
forest

Aspen
forest

Lodgepole
Whitebark
pine
Spruce-fir
pine
forest
forest
woodland

TREES
Aspen

Populus tremuloides

Douglas-fir

Pseudotsuga menziesii

Engelmann spruce

Picea engelmannii

Limber pine

Pinus flexilis

Lodgepole pine

Pinus contorta var.


latifolia

Pinus ponderosa

Pinus albicaulis

Abies lasiocarpa

Antelope bitterbrush

Purshia tridentata

Shinyleaf spiraea

Spiraea betulifolia var.


lucida

Chokecherry

Prunus virginiana

Ponderosa pine
Whitebark pine

Subalpine fir
SHRUBS

Common juniper

Juniperus communis

Gooseberry currant

Ribes montigenum

Grouse whortleberry

Vaccinium scoparium

Ninebark

Physocarpus spp.

Oregon boxleaf

Paxistima myrsinites

Rocky Mountain
juniper

Juniperus scopulorum

Rose

Rosa spp.

Russet buffaloberry

Shepherdia canadensis

Saskatoon
serviceberry

Amelanchier alnifolia

Skunkbush sumac

Rhus trilobata

Utah snowberry

Symphoricarpos oreophilus
var. utahensis

Wax currant

Ribes cereum

Bluebunch
wheatgrass

Pseudoroegneria spicata

Pinegrass

Calamagrostis rubescens

Idaho fescue

Festuca idahoensis

GRASSES

Limber
Ponderosa
pine
pine
Douglas-fir
woodland
forest
forest

Aspen
forest

Lodgepole
Whitebark
pine
Spruce-fir
pine
forest
forest
woodland

Common name

Latin name

Spike fescue

Leucopoa kingii

Mountain brome

Bromus marginatus

Wheelers bluegrass

Poa wheeleri

Geyers (elk) sedge

Carex geyeri

Ross sedge

Carex rossii

Arrowleaf
balsamroot

Balsamorrhiza sagittata

Bedstraw

Galium spp.

Fireweed

Chamerion angustifolium

Heartleaf arnica

Arnica cordifolia

Horsetail

Equisetum spp.

SEDGES

FORBS

Kinnikinnick

Arctostaphylos uva-ursi

Longstalk clover

Trifolium longipes

Meadow-rue

Thalictrum spp.

Nevada pea

Lathyrus lanszwertii

Creeping barberry
(Oregon-grape)

Mahonia repens

Wintergreen

Pyrola spp.

Sidebells
wintergreen

Orthilia secunda

Red baneberry

Actaea rubra

Sickletop lousewort

Pedicularis racemosa

Silky lupine

Lupinus sericeus

Sticky purple
geranium

Geranium viscosissimum

Timber milkvetch

Astragalus miser

Twinflower

Linnaea borealis

Western coneflower

Rudbeckia occidentalis

A dash indicates that the plant is absent or uncommon.


Whitebark pine occurs in northwestern Wyoming and northward in the Rocky Mountains.

178 Foothills and Mountains

The density of trees and understory plants in forests


varies considerably from place to place and also fluctuates through time.15 Where tree density is high, there
may be hardly any understory herbs or shrubs because
of too much competition from the trees for light, water,
and nutrients; in more open forests, the abundance
and diversity of smaller plants are often much greater.16
Lower tree density is found on sites that are too dry or
infertile to support a dense tree canopy, or in forests that
have been thinned by disturbances, as discussed below.
Snowdrifts persisting until early July restrict the growth
of some understory species, probably because the growing season is too short after the snow disappears.17
Plant parasites reduce tree growth, produce deformed
trees, or even kill some treesall problems in a setting
where wood production is a primary objective. Otherwise, native parasites add diversity to the forest community. Two that are especially important on Rocky
Mountain trees are dwarf mistletoe and comandra blister rust, both of which have coexisted with their hosts
for millennia.
Dwarf mistletoe (Arceuthobium americanum and other
species) is a flowering plant that is an obligate parasite
on pines (fig. 11.3). The pale green leaves and stems
are capable of some photosynthesis, but the plants
extend their roots into the hosts inner bark (phloem)

Fig. 11.3. Dwarf mistletoe is a parasitic flowering plant common on lodgepole and limber pine. The seed is spread to other
trees by explosive fruits, propelled by hydrostatic pressure.

and sapwood, where they obtain energy in the form


of carbohydrates, as well as water and nutrients.18 Tree

cides are avoided, because they generally have not been

growth is slowed, and sometimes the trees die.19 How-

effective, are expensive to apply over large areas, and

ever, bird species richness can be higher with mistletoe

have adverse secondary effects. Harvesting obviously

infection, suggesting that mistletoe eradication may be

infected trees is the most common approach, but often

undesirable where wood production is a less important

young mistletoe plants are difficult to see. Where foster-

objective.

ing the growth of a new generation of trees is a prior-

Mistletoe seed dispersal is accomplished by the ejec-

ity, clearcutting is the preferred option, whereby nearly

tion of seeds from the fruits by hydrostatic pressure,

all trees are cut, whether they are infected or not. This

with the seeds often traveling 2040 feet before landing

helps ensure that no infected trees are left behind. Most

on the leaves of another tree. The seeds have a slippery

likely, the prehistoric control of dwarf mistletoe was

coating and slide to the leaf base, where they germinate

periodic forest fires. Mistletoe now is probably more

in the spring. Infection occurs when the root success-

widespread than ever before because of fire exclusion,

fully penetrates the bark of 1- to 3-year-old twigs. The

and also because some early logging practices removed

mistletoe plant is hardly visible for several years, and

only the healthy trees.

new seeds are not produced for 56 years. Both healthy

Comandra

blister

rust

(Cronartium

comandrae),

and stressed pine trees are susceptible to mistletoe

another common native plant parasite, is a fascinat-

infection.

ing fungus that requires three very different plant spe-

Forest managers sometimes work to reduce the abun-

cies in close proximity to accomplish its life cyclein

dance of mistletoe to maximize tree growth. Herbi

this case, lodgepole pine, big sagebrush (or a few other

Mountain Forests

species), and a small herbaceous plant known as comandra or bastard toadflax. Comandra, the plant, is commonly found growing in dry mountain meadows and
is itself an obligate root parasite on sagebrush and various other plants. The disease therefore usually develops
only where sagebrush, comandra, and lodgepole pine
occur near one another. This arrangement is common
and the disease is widespread.20 Tree death cannot usually be attributed to the rust directly, but the treetops
die back and often a forked trunk develops. This is a
problem for forest managers focused on wood production, because it causes a reduction in the amount of harvestable wood. As with dwarf mistletoe, however, it is a
native species and is part of the ecosystem.
Animals also have significant effects on forest communities (table 11.2). For example, deer and elk can
reduce the rate of aspen regeneration by browsing the
small sprouts produced by the root system, and seedcaching birds (such as Clarks nutcracker) influence the
distribution of limber and whitebark pine in forests
as well as the foothills (see chapters 10 and 15). Pine
seeds are an important source of food for red squirrels
(fig. 11.4), with the cones being cached by the squirrels for winter consumption. However, not all seeds are
eaten, partly because the pine trees have adaptations
that deter squirrels; for example, the cones are resinous
and have thick scales. 21 All coniferous forests in Wyoming have red squirrels, but where they are not found,
to the north and west of Wyoming, the cones are less
massive and have twice as many seeds. Also, the seeds
are more easily removed. In such situations, the nutcrackers are more abundant. Where the nutcrackers
and squirrels do coexist, Wyoming ornithologist Craig
Benkman and his associates have found that the birds
bill is longer and stouter, presumably to give the bird a
better chance of prying open the scales of the heavyduty cones that have evolved in the presence of the
squirrels. 22
The serotinous cones produced by lodgepole pine
retain large numbers of seeds, which has made the
cones an especially rich food source for squirrels. As

Table 11.2. Some mammals, birds, amphibians, and


reptiles found primarily in mountain landscapes
MAMMALS
Bear, black
Bear, grizzlya
Bobcatb
Chipmunk, leastb
Cougar
Coyoteb
Deer, muleb
Elkb
Ground squirrel, Wyomingb
Hare, snowshoe
Marmot, yellow-bellied
Marten, American
BIRDS
Bluebird, mountainb
Chickadee, black-cappedb
Chickadee, mountain
Crossbill, red
Crossbill, white-winged
Dipper, American
Duck (various species)b
Finch, Cassins
Flicker, northernb
Flycatchers (various species)b
Goshawk, northern
Grosbeak, evening
Grosbeak, pine
Grouse, dusky
Grouse, ruffed
Hawk, red-tailedb
Jay, gray
Jay, pinyonb
Jay, Stellers
Junco, dark-eyedb
Kinglet, ruby-crowned
Nutcracker, Clarks
Nuthatch, red-breasted

Owl, boreal
Owl, great hornedb
Pipit, American
Raven, commonb
Robin, Americanb
Rosy-finch, gray-crowned
Rosy-finch, black
Rosy-finch, brown-capped
Sapsucker, red-naped
Sapsucker, Williamsons
Solitaire, Townsends
Sparrow (various species)b
Swallow, violet-green
Tanager, western
Thrush, hermit
Thrush, Swainsons
Veery
Warbler (various species)b
Woodpecker, American
three-toed
Woodpecker, black-backed
Woodpecker, downyb
Woodpecker, hairyb

AMPHIBIANS
Toad, boreal
Frog, boreal chorusb
Frog, Columbia spotted

noted, however, thicker cone scales have evolved to

REPTILES

minimize accessibility. This, in turn, has favored the

Boa, rubber
Gartersnake, wanderingb

evolution of larger jaw muscles in squirrels (compared

Moose
Mouse, deerb
Pika, American (alpine)
Pocket gopher, northernb
Porcupineb
Sheep, bighorn
Squirrel, red
Vole, heather
Vole, southern red-backed
Wolfa
Wolverine
Woodrat, bushy-tailedb

to those of squirrels in Douglas-fir forests, which do

not have serotinous cones). Craig Benkman and Matt

Frog, northern leopardb


Frog, wood
Salamander, tigerb

Snake, smooth green

Found only in the Greater Yellowstone Ecosystem.


Also found away from the mountains.

179

180 Foothills and Mountains


Fig. 11.4. Red squirrels have coexisted
with lodgepole pine for millennia.
Populations that depend on serotinous
lodgepole pine cones have extraordinarily strong jaw muscles for extracting the nutritious seeds from the
hard woody cones. The squirrels are
omnivores, as they also eat mushrooms,
bird eggs, and nestlings when available.
Photo by Fred Walsh.

Talluto concluded that seed predation by red squirrels

into forests dominated by ponderosa pine or Douglas-fir

reduces the occurrence of trees with serotinous cones.

at lower elevations, lodgepole pine at mid-elevations,

Thus, the characteristics of serotinous cones and the

and Engelmann spruce and subalpine fir at higher eleva-

proportion of serotinous trees in a specific forest is the

tions. 24 Considerable overlap occurs in the elevational

result of natural selection exerted by both the squirrels

distribution of these tree species, and in some areas the

and fire.23

foothill vegetation grades directly into lodgepole pine

Forest communities include hundreds or thousands

or Douglas-fir, as, for example, on the east slopes of the

of species of native insects. Most are inconspicuous

Medicine Bow Mountains and around Jackson Hole and

and spend their lives quietly pollinating flowers or

the Sunlight Basin. Ponderosa pine is rarely found in

decomposing dead organic matter. A few insects attract

western Wyoming, where Douglas-fir forests usually

attention, however, because they feed on living trees,

border the foothill vegetation.

sometimes killing their hosts. One groupthe bark

Intermingled with the coniferous forests are moun-

beetleshas gained prominence recently because these

tain meadows, also known as parks, along with rib-

insects are changing many forests in the Rocky Moun-

bon forests, snowglades, aspen groves, various kinds of

tains and westward. The causes and effects of bark

riparian communities, and shrublands or woodlands

beetle outbreaks are examined later in this chapter.

dominated by mountain big sagebrush and limber pine.


The upper treelinewhere the mountains extend that

The Forest Mosaic: Variation with Elevation


and Topography

highis characterized by short, wind-pruned Engelmann spruce, subalpine fir, and limber pine in most of
Wyoming ranges, though whitebark pine is typical of

The causes of vegetation changes with elevation have

the treeline in northwestern Wyoming and the North-

been of interest to scientists for more than two cen

ern Rockies. In central Colorado, bristlecone pine can

turies. In the Rocky Mountain region, foothill grass-

be found at upper treeline and also forms subalpine

lands, shrublands, and woodlands commonly grade

woodlands in some areas.

Mountain Forests

Disturbance and Change over Time: Another


Important Source of Variation in Forest
Communities
Forests are always changing, sometimes so slowly that
the changes are imperceptible. The ecological processes
that drive slow changes include competition among
trees for light and other resources, selective feeding

disturbances in forests are especially noticeable be


cause they commonly occur over large areas and their
impacts often last for decades (fig. 11.5). 26 In any mountain landscape, a shifting mosaic of forests responds
both to local environmental conditions and to the
effects of disturbances, some of which were recent and
others centuries old.

by herbivores, tree death, seed dispersal, and germination. Other changes are rapid and obvious, usually
resulting from discrete disturbances that kill or injure
many of the trees and other forest organisms. Examples include fire, an insect outbreak, a severe windstorm, and timber harvesting. Forests also change in
response to variation in the local climate. For example,
a severe but temporary drought may injure or kill susceptible plants; or a long-term directional change in
climate may lead to a gradual change in overall species
composition, with new species moving in while others
decline. In addition to the adaptations already discussed, surviving in Rocky Mountain forests requires
an ability to bounce back after the disturbances that
have occurred for millennia. 25
Landscape changes caused by disturbances and bi
ological processes are superimposed on the patterns
that develop in response to environmental gradients
in local climate and soils. Both kinds of pattern are
important in all vegetation types, but the effects of

Forest Fires
Two key aspects of all disturbances are their frequency
and severity. The natural frequency of fire varies from
decades to centuries among different forest types. Forest
fires tend to be most frequent in the foothills and lower
elevations of the mountains, where precipitation is sufficient to support flammable vegetation and where summers are often dry enough to permit fires to ignite and
spread over large areas. Late nineteenth- and twentiethcentury fire exclusion reduced fire frequency in many
lower-elevation landscapes.27 Fires became less frequent
in some areas because livestock grazing reduced the
abundance of fine-textured fuels, and roads, fields, and
towns broke up formerly continuous expanses of flammable vegetation.
At higher elevations, fires are notably less frequent
because of the typically moist conditions in those en
vironments: the snowpack melts later in the spring,
or not until midsummer. Also, summer rainstorms
often dampen the fuels, and relative humidity is often

Fig. 11.5.Tree-killing bark beetles of the genus Dendroctonus,


such as this mountain pine beetle, are native insects that have
coexisted with pine trees for thousands of years. The adults
are about 0.2 inch in length, though their wingspan is much
longer. Many species of bark beetles do not kill trees, subsisting instead on the inner bark of trees that have recently died
from other causes. Photo by Dion Manastyrski,from the collection of Lorraine Maclauchlan.

higher than at lower elevations. Lightning ignites fires


every year in high-elevation forests, but most fail to
spread beyond the ignition point because fuels are too
wet to burn. Only in unusually dry years are weather
conditions suitable for extensive fire spread. Significantly, its only during those historically infrequent
years that large fires occur in high-elevation forests.
Compared to climatic factors, fire suppression has
had a small impact on the frequency and size of highelevation fires.
Two related terms are fire intensity and fire severity. Fire intensity refers to the energy released in a fire,
expressed as BTUs per minute per unit area; fire severity refers to the effects of that energy release on organisms or ecosystems. Intensity must be measured while
the fire is in progress, which is difficult. Severity, however, can be estimated after a fire by characterizing tree

181

182 Foothills and Mountains

mortality and combustion of organic matter. Severity


ranges from surface fires that only consume a portion
of the forest floor and kill small trees, without harming most larger trees, to high-severity stand-replacing
crown fires that kill most of the trees and consume
essentially all of the forest floor. Forest fires typically
burn in a highly heterogeneous manner, responding
to wind direction, slope and aspect, and variation in
local weather and fuel conditions. The result is a mosaic
of severely burned patches, light or moderately burned
patches, and unburned forest.28
Each tree species has adaptations that enable it to persist in the face of periodic disturbance, and each major
forest type tends to be characterized by a different kind,

Bark Beetle Outbreaks


Rapid increases in the populations of some insect species
can lead to major disturbances that change forest structure and dynamics.31 The mountain pine beetle is the
insect most often mentioned as a disturbance agent in
Rocky Mountain forests, but other species of bark beetle
are also involved, each specializing in feeding on a particular tree species or group of species. These include the
Douglas-fir beetle, spruce beetle, and western balsam bark
beetle (which kills fir trees). All are natives. Defoliating
insects, notably the western spruce budworm and aspen
tent caterpillar, also modify forests. Because bark beetles
have had such dramatic effects over such extensive areas

frequency, and severity of disturbance. For example,


the thick bark of mature ponderosa pine and Douglasfir trees protects the critical inner bark tissues from heat
injury and enables these trees to survive low-intensity
surface fireswhich historically were the most common type of fire in many low-elevation ponderosa pine
and Douglas-fir forests. However, all conifers are killed
by fires that burn through the crowns of the trees, a
common type of fire in montane and subalpine forests.
Lodgepole pine is well adapted to such high-severity
fires, because even though the adult trees are usually
killed, each tree often has hundreds of cones that store
large numbers of viable seeds. At least some of the seeds
survive even an intense fire, as discussed later in this
chapter. These seeds germinate and grow best on the
bare soil exposed by a severe fire. Aspen also persist in
burned forests, despite fire killing all the aboveground
portions of the tree, because its roots usually survive
and produce abundant new sprouts that grow rapidly
into new trees. Soil is an excellent insulating material
for the roots.
Fires also affect animal abundance and distribution
patterns. For example, three-toed and black-backed
woodpeckers typically nest and reproduce in coniferous forests that have recently burned or been affected
by bark beetle outbreaks, because the insects inhabiting
the dead or dying trees provide an abundant and readily available food source. 29 Ungulates and bears benefit
from the greater abundance of herbaceous or shrubby
plants after such disturbances, and raptors are drawn
to disturbed forests because rodents and other prey are
more visible.30

Fig. 11.6.Numerous spots of resin and yellow sawdust on


the bark indicate that this lodgepole pine has been invaded
successfully by mountain pine beetles. Each spot represents
the entrance of one beetle. The beetles introduce the bluestain fungus that soon plugs the sapwood, causing tree death
within a few months. The leaves are reddish-brown by the
following summer (as illustrated in figs. 16.4 and 16.9 for
infested ponderosa pine). This pine also has a scar on the
trunk, possibly caused by a porcupine if not a fire.

Mountain Forests

Fig. 11.7. Many lodgepole pine on the mountain slope in the


background have been killed by mountain pine beetles. In
other areas, Engelmann spruce forests are affected similarly
by the spruce beetle. This area is near Green River Lakes in
the Wind River Mountains. Large numbers of elk are attracted
to this area in the winter because of a winter feeding program,

which may account for the poor condition of the aspen in


the foreground (see chapters 10 and 15). The soil on this
south slope is relatively moist, as indicated by the presence
of shrubby cinquefoil, mountain silver sagebrush, common
juniper, and western snowberry. The small conifer next to the
aspen is limber pine.

in recent decades, they are the focus of this section. Only

In late July and August the female beetles emerge

a few of the hundreds of different kinds of bark beetles in

from the bark and fly toward other trees. They are most

western North America actually kill trees; more than 99

often attracted to larger trees (those greater than 68

percent feed on trees that have already died from other

inches in diameter), which surely is adaptive because

causes, contributing to the decomposition of bark and

such trees provide the beetle larvae with more food,

woodan important ecosystem service.32

namely, the phloem, cambium, and sapwood of the

All tree-killing bark beetles have similar life histo-

inner bark. In addition to visual clues, the beetles can

ries and ecological effects, which can be illustrated by

be attracted by the odors emitted by stressed or weak-

the mountain pine beetle. This insect attacks all native

ened trees, whether caused by drought, age, scarring by

pines in Wyoming (lodgepole, ponderosa, whitebark,

disturbances, or other factors.34

and limber). 33 In most years the beetle population is

After landing, the female beetle bores a hole through

low, surviving in weakened or recently fallen trees,

the bark and is soon cutting egg-laying galleries in the

with populations that are referred to as endemic. Peri-

outer sapwood and phloem. However, if the tree is in

odically, however, the population increases rapidly, and

good health, resin is produced in such quantities that

vast numbers of trees are killedan event referred to as

the beetle is pitched out before any damage to the tree

an outbreak, epidemic, or eruption (figs. 11.6 and 11.7).

occurs. Such trees are easily recognized by small globs

An understanding of the life cycle of the beetle and of

of resin adhering to the external surface of the bark,

the causes of its population explosions is fundamental

in which a dead beetle can often be found. Healthy

to understanding forest ecosystem dynamics.

trees also produce chemicals that are directly toxic to

183

184 Foothills and Mountains

the beetles. Less vigorous trees are unable to produce

temperature threshold applies in all places and at all

enough resin or chemical defenses to repel the beetles.35

times. Great genetic and physiological variations occur

Following a successful attack, one often sees fine saw-

among different beetle speciesand among geographi-

dust around the holes made by the beetles and at the

cally segregated populations of the same speciesin

base of the tree.

the rate, timing, and effectiveness of cold hardening.

The mountain pine beetle is a tiny insect about 0.2

In general, however, bark beetle larvae are most vulner-

inch long, about the size of a grain of rice. The first few

able to a spell of cold weather in late fall and early win-

female beetles to successfully burrow through the bark

ter, before the cold hardening process is completed, as

promote mass attacks by emitting an aromatic chemi-

well as in late winter and early spring, when warmer

cal, known as an aggregating pheromone, which attracts

temperatures have stimulated the beginning of a break-

Curi-

down of the antifreeze compounds. In the late spring or

ously, once a sufficient number of beetles have been

summer, after further feeding and maturation, another

attracted to the tree, the successful beetles emit a dis-

generation of adult beetles emerges to attack other live

aggregating pheromone, which repels other beetlesan

trees, and the cycle continues.

other mountain pine beetles to the same tree.

36

amazing adaptation that helps ensure the survival of

This life cycle enables bark beetles to persist at low

the species. If too many beetles attack the same tree,

numbers, killing the occasional weak or stressed tree

fewer would reproduce successfully. Such specific

but having little impact on overall forest structure.

responses to specific stimuli by both the beetles and the

How, then, does a bark beetle population suddenly

tree are a reflection of the long history of co-evolution

explode into an outbreak in which the beetles kill so


many trees that everyone notices the effects? Two con-

of the insect and its host. 37


Following successful attack, the females lay eggs in

ditions are required: suitable climate conditions and

their galleries. At the same time, the spores of a fungus

susceptible trees. Warmer temperatures affect both, by

known as the blue-stain fungusare introduced.

The

enhancing survival of beetle larvae and contributing

fungal filaments grow into the sapwood, where they

to the development of tree water stress, thereby reduc-

restrict the flow of water and nutrients and kill the

ing the amount of resin produced and making the trees

tree in less than a year.39 The beetle eggs soon hatch,

more susceptible to beetle attack. Ecologist Teresa Chap-

and the larvae begin feeding on the inner bark, mov-

man and colleagues found that recent mountain pine

ing horizontally around the tree, cutting through the

beetle outbreaks in northern Colorado and southern

phloem and stopping the flow of carbohydrates from

Wyoming closely followed a period of warming tem-

the leaves to the roots. Although this girdling surely con-

peratures, which included several years of reduced pre-

tributes to the death of the tree, trees stressed only by

cipitation.42 Large numbers of trees were vulnerable to

girdling can live for 45 years.

beetle attack, because they were stressed by drought and

40

38

It seems clear that the

plugging of the sapwood by the blue-stain fungus is the

were of sizes preferred by beetles.

primary factor causing tree death. Notably, the larvae

As forests age, the trees experience increased compe-

also obtain important nutrients by feeding on the blue-

tition from their neighbors and many become more sus-

stain fungus.41

ceptible to beetle epidemics. Defensive chemicals, such

In the fall, the larvae begin to produce antifreeze-

as resin, are thought to have low priority in the alloca-

like chemicals in their bodies. They also eliminate

tion of limited carbon compounds by trees, compared

substances that could serve as ice-nucleating agents.

to the maintenance of leaf area and young roots, with

This process of cold hardening greatly increases winter

the result that less resin is produced and beetle inva-

survival with minimal protection from the coldonly

sion is more likely. Also, damaged trees, such as those

a thin layer of bark. Nevertheless, long periods of very

scarred by lightning strikes, fire, or logging activity,

cold temperatures at the right time of year can result in

are more susceptible. Some research suggests that the

high mortality of beetle larvae, and such cold-related

microclimate of closed, old-growth forests is a contrib-

mortality may be an important mechanism preventing

uting factor to initiating an outbreak. According to this

or terminating a bark beetle outbreak. No single lethal

hypothesis, closed canopies minimize air movement,

Mountain Forests

causing beetle-attracting odors to concentrate. Beetles

With less competition for light, water, and nutrients

are then attracted to such forests from a larger area, and

following the death of overstory trees, the understory

with high beetle densities, even the more vigorous trees

trees often grow much faster than they did before the

are susceptible to invasion.43 Under such conditions,

outbreak. For example, height growth rate of many sur-

beetle populations increase rapidly in one year. If the

viving lodgepole pine and subalpine fir trees doubled

forest is opened up by whatever means, the odor plume

after the mountain pine beetle outbreak in a northern

is dispersed and becomes less effective. Beetle parasites

Colorado study area, and radial growth rate increased

and beetle predation by birds may also be factors that

two- to threefold after a mountain pine beetle out-

reduce bark beetle populations.

break in the Greater Yellowstone Ecosystem.48 Modeling

Once an extensive outbreak is under way, the thou-

studies suggest that this rapid growth of the surviving

sands of adult beetles emerging from infested trees can

understory can restore pre-outbreak levels of basal area

overcome the defenses of healthy treeseven some rel-

within about 80 years.49 Recovery of basal area may be

44

atively small onesas well as those of unhealthy trees.

even faster under some conditions, as observed after a

The healthy hosts provide a superior food resource for

mountain pine beetle outbreak in Colorado in the 1980s.

the beetle larvae; hence, even more new adults emerge

Ninety percent of the basal area that had been present

the following year. When conditions are especially

prior to the outbreak had regrown after only 30 years in

favorable for the beetles, the outbreak may be so exten-

some lodgepole pine stands.50 In forests where lodgepole

sive that it becomes a regional or even continental phe-

pine dominated the pre-outbreak overstory but subal-

nomenon, as occurred in western North America during

pine fir and Engelmann spruce were more abundant in

the late 1990s and early 2000s.45 An outbreak usually

the understory, the composition of the post-beetle forest

ends only after a cold winter kills many larvae or when

shifts toward more fir and spruce; however, where lodge-

susceptible trees are no longer availableeither because

pole pine dominated both the overstory and the under-

the conditions causing host-tree stress have been ameli

story, lodgepole continues as the dominant species.51

orated in the area or because the beetles have simply


killed so many trees of suitable size. The last mountain
pine beetle outbreak in northern Colorado and south-

Major Forest Types

ern Wyoming apparently subsided because the beetles

Rocky Mountain forests are easily classified into sev-

could no longer find enough suitable trees; temperatures cold enough to terminate the outbreak were not
recorded.
An important feature of the disturbances caused
by most bark beetle species is that nearly all mortality
occurs in the larger trees, namely, those greater than
about 68 inches in diameter. Most smaller trees and
saplings usually survive, even where the beetles kill
most of the large trees in the overstory. For example,
severe mountain pine beetle outbreaks in lodgepole
pine forests of southeastern Wyoming and northern
Colorado during the 1990s and 2000s reduced the canopy by 10 to more than 70 percent in many stands.46
Nevertheless, even in the most severely affected stands,
the smaller trees that were not attacked were generally

eral major types (see table 11.1), each of which can be


further subdivided. The most thorough classification
was done using Rexford Daubenmires habitat-type
approach, where forests are classified according to two
factors: (1)the tree species that would dominate a stand
after a very long time without disturbance, when the
vegetation has presumably equilibrated with prevailing climate and soils, and (2) one or more understory
plant species that are indicators of other environmental
conditions. This classification assists in evaluating sites
for wood production, wildlife habitat, and other forest
values.52

Ponderosa Pine Forest

numerous enough to meet or exceed U.S. Forest Service

Ponderosa pine is the most widespread conifer in the

requirements for stand regeneration. Also, less competi-

western United States and is a major tree species at lower

tion from older trees enables the establishment of new

elevations throughout its range. In Wyoming it grows

tree seedlings.47

primarily in the eastern portions of the state, where

185

186 Foothills and Mountains

Fig. 11.8.On the left is an open ponderosa pine forest on


a moderately moist site in the Laramie Mountains south of
Douglas. The thick bark of mature ponderosa pine protects
many trees from easily controlled surface fires that kill
smaller trees, although some are scarred (see fig. 11.9). In the

same area, ponderosa pine forests develop extremely high tree


densities after long periods without fire or other disturbances
that coincide with climatic conditions favorable for seedling
establishment, as illustrated on the right. Such dense forests
are highly susceptible to severe wildfires. Elevation 6,700 feet.

summer precipitation is higher and the growing season

The importance of warmer temperatures was suggested

is warmer and longer (see figs. 1.5 and 3.4), especially

by a laboratory study that found that ponderosa pine

in the Black Hills and on the east slopes of the Bighorn

seedlings are more sensitive to cold temperatures than

and Laramie mountains. The importance of summer

were those of lodgepole pine, which is found at higher

precipitation is illustrated by the absence of ponderosa

elevations. 53 Probably for the same reason, ponderosa

pine on the southern end of the Bighorn Canyon, in the

pine is generally absent from the frost pockets in depres-

desert-like Bighorn Basin, and its presence at the same

sions, where cold air accumulates. Ponderosa pine is

elevation on the northern end of the canyon, where

a relative newcomer near the northeastern limit of its

summer rainfall is much higher (19 inches compared

distribution, such as in Wyoming, and probably is still

to 7 inches). Ponderosa pine is also abundant in por-

moving northward from Pleistocene refugia far to the

tions of the Uinta Mountains in Utah, the east slopes of

south (see chapter 2). Climate warming may accelerate

the Colorado Front Range, and throughout the moun-

this migration.54

tains of southern Colorado, New Mexico, and Arizona.

Tree density in ponderosa pine forests is a function

In all these areas, as on the Great Plains, a large por-

of local growing conditions and disturbance frequency

tion of the annual precipitation occurs in the summer.

and severity (fig. 11.8). On relatively wet sites, ponder-

Mountain Forests

osa pine can form a nearly closed canopy with a sparse


understory; on drier sites the forest is more open, or even
savanna-like, and often has a well-developed understory of shrubs, grasses, and other herbaceous plants.
Open ponderosa pine stands can also be maintained
by frequent surface fires that kill young seedlings and
saplings. The larger, older trees usually are not killed by
surface fires, owing to their thick bark and high crowns,
but they are often scarred by the fire. The scars can be
analyzed using dendrochronology to reconstruct local
fire history over the past several centuries (fig. 11.9).
Such fires burn the tops of understory shrubs and herbaceous plants, but the roots survive and soon produce
new sprouts, a type of vegetative reproduction that is
lacking in the conifers. Mountain pine beetle outbreaks
also thin ponderosa pine stands, although the resulting stand structure is different from that created by fire,
simply because the beetles tend to kill more of the large
trees and fewer of the saplings.
At the scale of an individual ponderosa pine stand,
fires commonly recurred at intervals of a few years to
a few decades prior to disruption of the historical fire
regime in the late 1800s.55 In some areas these fires were
predominantly low-severity surface burns that helped
maintain an open forest structure. Other ponderosa
pine forests were characterized by a variable-severity
fire regime that included a patchy mix of low- and
high-severity burning, both during individual fires and
in individual stands over time.56 High-severity (standreplacing) burning was an important component of the
fire regime primarily on wetter sitesduring dry years

Fig. 11.9. Ponderosa pine and Douglas-fir with thick bark


often survive surface fires, even though the base of the tree is
scarred. Fire scars develop when the heat scorches a portion
of the cambium, a layer of cells found on the inner side of
the bark that produces new bark and sapwood. Such scars can
be used to determine the years that fires occurred and the
number of years between successive fires. This photo of a ponderosa pine was taken in Devils Tower National Monument.

in the ponderosa pine zone, such as at higher elevations


and in mixed-conifer stands where other species, such
as lodgepole pine or Douglas-fir, were co-dominant. In

historically in the wetter northern part than in the drier

the Colorado Front Range, for example, fires at lower ele-

southern part.58

vations, especially near the prairie-forest ecotone, were

Severe burns were followed by either gradual or

predominantly low-severity surface burns that helped

rapid recovery of forest cover, depending on climatic

maintain low-density ponderosa pine forests. At mid-

conditions and seed availability. In some settings, a

elevations, ponderosa pine forests burned with variable

dense stand of ponderosa pine develops after a high-

severity, so that patches of stand-replacing fire were

severity fire; in other settings, such as on dry sites fol-

interspersed with areas of low-severity surface burn-

lowing severe burns, the post-fire stand may be sparse

ing. At the upper reaches of the ponderosa pine zone,

and slow to re-establish.59 Ponderosa pine seeds are rel-

where ponderosa transitions into lodgepole pine forest,

atively large and tend to disperse over relatively short

stand-replacing fires burned large areas of ponderosa

distances; consequently, the interior of a large patch of

and lodgepole pine forest.57 Similarly, in the Black Hills,

stand-replacing fire may become reforested very slowly,

dense stands and high-severity fires were more common

as has been seen after some recent fires in Colorado. The

187

188 Foothills and Mountains


Fig. 11.10. Douglas-fir forest
near Jenny Lake in Grand
Teton National Park. Elevation 6,850 feet.

establishment of new ponderosa pine seedlings tends to

Laramie Mountains near Casper and near Laramie Peak

be episodic, occurring primarily during periods of rela-

west of Wheatland, and a large fire of variable fire sever-

tively wet conditions, with or without fire.60 As a result,

ity burned near Laramie Peak in the summer of 2012.

a ponderosa pine forest often contains only three or

The same is true elsewhere in the region.

four easily identified age classes.


The earliest explorers surely saw a great diversity of
ponderosa pine forests as they traveled westward. How-

Douglas-fir Forest

ever, a combination of fire exclusion and heavy grazing,

Douglas-fir and ponderosa pine share several ecological

which enhances tree seedling establishment by reduc-

characteristics and sometimes they are found together.

ing competition from grasses and forbs, caused tree

Specifically, both species have a thick bark that often

density in many stands to increase during the twenti-

enables the trees to survive low-intensity surface fires.

eth century, especially on drier sites at low elevations

Also, the optimal environment for both is at lower eleva-

(see fig. 11.8). Wetter conditions also promoted seedling

tions, in the foothills and lower mountain slopes below

establishment and survival in areas that were logged in

about 8,500 feet (see fig. 11.1). Usually Douglas-fir for-

the early 1900s, when tree removal favored the growth

ests are found at slightly higher elevations than pon-

of young trees in many places.61 All these factors favored

derosa pine, or on somewhat wetter sites; for example,

the development of more extensive and uniform wood-

north-facing slopes often support relatively pure stands

lands than had occurred historically. Recent large fires,

of Douglas-fir, whereas on the opposite, south-facing

bark beetle outbreaks, and drought-caused tree mor-

slopes, stands may be dominated primarily by ponderosa

tality, as well as innovative new timber harvesting

pine. Limestone and other sedimentary substrata appear

methods, may be restoring some of the diversity that

to be especially favorable for Douglas-fir.62 Some research

characterized historical ponderosa pine forests. Nota-

suggests that Douglas-fir is more tolerant of competition

bly, since about 2000, extensive mountain pine beetle

for water, nutrients, and light in the understory than is

outbreaks have affected ponderosa pine forests in the

the pine, and that it can eventually replace the pine in

Mountain Forests

the absence of disturbance. In Wyoming, Douglas-fir is

matic conditions were very dry. For example, widespread

most common in the western half of the state, where it

fires in 1856 burned Douglas-fir forests in portions of

occurs just below the lodgepole pine forest zone. It is also

Jackson Hole, northern Yellowstone National Park, and

abundant on the east slopes of the Colorado Front Range,

the Centennial Valley just west of Yellowstone Park.64

and in adjacent Idaho and Montana, such as to the west


and north of Yellowstone National Park (fig. 11.10).

Insects can also cause variation in forests of Douglas-fir, especially the Douglas-fir beetle and the western

As with ponderosa pine forests, tree density and

spruce budworm (figs. 11.11 and 11.12). The bark beetles

understory cover in Douglas-fir forests vary with local

kill the larger trees, whereas the budworm kills mainly

environmental conditions and disturbance frequency

smaller individuals. Rarely do the insects kill all trees

and severity, ranging from sparse, savanna-like stands to

in a stand.65 Mixed-age Douglas-fir stands develop on

closed-canopy stands. Historical fire regimes were simi-

dry sites or where fires or insects have caused only par-

lar to those in ponderosa pine forests, with fire intervals

tial mortality of the canopy; even-aged stands (that is,

perhaps somewhat longer and high-severity fire perhaps

those made up of trees that are roughly the same age)

more common in Douglas-fir, especially on wetter sites.63

also develop, especially after high-severity fires or other

Large stand-replacing fires occurred in years when cli-

disturbances that kill all the trees.66

Fig. 11.11. Douglas-fir woodland in northern Yellowstone


National Park, 15 years after a severe outbreak of Douglas-fir
bark beetle. The beetles killed most of the larger trees in the
stand, which stimulated the growth of small trees, shrubs,
and herbaceous plants. Notably, in contrast to lodgepole pine,
Douglas-fir snags tend to break off rather than uproot. Photo
by Dan Donato.

Fig. 11.12. Larvae of western spruce budworm, a moth, commonly attack the buds and young needles of Douglas-fir, as seen
here in northern Yellowstone National Park in 2012. This budworm also attacks white fir in Colorado, Utah, and Idaho, and,
to a lesser extent, Engelmann spruce. It is said to be the most
widespread forest defoliator in western North America (Leatherman et al. 2009). Tree growth is slowed, but unlike bark beetles,
the western spruce budworm usually does not kill its host.

189

190 Foothills and Mountains

Fig. 11.13. An even-aged stand of 150-year-old lodgepole pine


above Centennial in the Medicine Bow Mountains. Tree-ring
data suggest that this forest was initiated following a fire. The
forest understory is dominated by dwarf huckleberry. Clusters

of small lodgepole pine are developing in openings. The blue


mark on the tree indicates that this tree will soon be cut,
probably for safety reasons. Elevation 8,400 feet.

As with ponderosa pine forests, tree density in many

Unlike the other conifers, lodgepole pine thrives on

Douglas-fir stands increased during the twentieth cen-

nutrient-poor soils (such as those that develop on gran-

tury as a consequence of fire suppression, grazing, and

ites and rhyolites) and is less common on more fertile

climatic conditions favorable for tree establishment. In

limestone-derived soils.68

some places, especially around meadows with big sage-

Lodgepole, as it is commonly called, has been viewed

brush, one can see the skeletons of sagebrush that died

as a pioneer species after fires or other severe distur-

for lack of sunlight as the Douglas-fir expanded into the

bances, because it generally does not tolerate the forest

shrubland. Recent fires and insect outbreaks also diver-

understory environment and is often the first tree to

sify stand structures.

become established after a disturbance, especially where


the soils are relatively dry and infertile. Under such con-

Lodgepole Pine Forest

ditions, lodgepole forms what is known as an even-aged


stand of trees. In fact, tree ages may differ by 20 years

The most common tree in the mountain forests of

or more, but the forest clearly became established after

northern Colorado, Wyoming, and much of the North-

a widespread disturbance. As the forests age and new

ern Rockies is lodgepole pine (fig. 11.13). It usually grows

gaps in the canopy are createdby localized windthrow

at mid-elevationsabout 6,00010,500 feet in north-

or other low-severity disturbanceslodgepole seedlings

ern Wyoming and about 7,00011,500 feet in southern

can become established and grow into the canopy.

Wyoming.

This pine is found at higher elevations on

In this way, tree establishment becomes episodic and

drier sites or in areas that have burned in the past cen-

patchy, resulting in a forest with multiple age classes.

tury or two. Engelmann spruce and subalpine fir may

The probability of lodgepole seedling establishment in

be co-dominant with lodgepole on relatively moist sites.

the forest understory appears to increase with moisture

67

Mountain Forests

Fig. 11.14. Lodgepole pine forests along the Rockefeller


Parkway between Yellowstone and Grand Teton national
parks. The stand on the left originated after the fires of 1988,
24 years before this photo was taken. The young trees are
very dense, because many of the pre-fire trees had serotinous cones, and also because the conditions for seedling

establishment during the two years after the fire were


favorable. The stand on the right has not burned for more
than 250 years. Many of the dead trees visible in the older
stand have been killed by mountain pine beetles during the
past 10 years. A riparian meadow and shrubland are in the
foreground.

availability, even without a canopy gap, but so does the

closed and on the tree for many years until higher than

chance of spruce and fir establishment.

usual heat causes the cone scales to open, such as dur-

Usually fires in lodgepole pine forests are infrequent,

ing a forest fire.70 The cones also open after falling to

separated by many decades or centuries in any given

the forest floor if they are exposed to the sun for a long

stand. This is primarily because of the cool, relatively

enough period.71 Numerous cone crops accumulate in

moist environments inhabited by lodgepole pine,

the treetops as the forest develops. Most of the seroti-

where snow persists until late spring and frequent sum-

nous cones open following a fire, and the seed is dis-

mer rain showers commonly keep fuels too wet to sup-

persed by wind at a time when a mineral soil seedbed

port extensive fire in most years. Only in exceptionally

and low competition for resources help to ensure seed-

dry years do large fires occur in lodgepole pine forests.

ling survival. At such times, more than a hundred thou-

Under these conditions, fires tend to be severe and

sand seeds may be dispersed per acre in a single year.72

stand-replacing, often burning over large areas.

69

Interestingly, not all lodgepole pine trees produce

In contrast to ponderosa pine and Douglas-fir, lodge-

serotinous cones. Individual trees tend to have either all

pole pine trees have thin bark and are easily killed by

serotinous or all nonserotinous cones, though a few trees

even low-intensity fires. However, lodgepole is well

may produce both types.73 The proportion of serotinous

adapted as a species, if not as individuals, to a high-

trees in stands of lodgepole ranges from 0 to nearly 100

severity fire regime because of its ability to produce

percent, with the highest proportions in young stands

serotinous cones (figs. 11.14 and 11.15). These cones are

at lower elevations, where the last major disturbance

produced more or less continuously, but they remain

was fire. The lowest proportions are in older stands at

191

192 Foothills and Mountains

Fig. 11.15.On the left, serotinous cones on a lodgepole pine


branch that has recently fallen to the ground. The diameter of
the branch indicates that the cones have remained closed for
many years, accumulating seed on the tree. One of the cones
is beginning to open, most likely because of higher temperatures at the soil surface than in the trees canopy. Cones that

remain on the tree typically open with the heat of a fire, as


shown in the center photo. The pine needles have burned,
but only the exterior of the cone scales was blackened. In
contrast, on the right, the green leaves around the open cone
indicate that the heat of a fire was not required for seed dispersal from the cones of this tree.

higher elevations, where the last major disturbance was

such as sagebrush-dominated shrublands. Some have

a windstorm, insect outbreak, or other nonfire distur-

suggested that forest management, in both wildlands

bance.74 To some degree, this pattern is genetically con-

and timber production areas, should allow for both fire

trolled and reflects long-term differences in selection

and nonfire disturbances, so that both genotypes can

pressures in different environments. Because fire inter-

be sustained, thereby maintaining biological diversity.

vals generally become longer with increasing elevation,

Seeds in both serotinous and nonserotinous cones

reaching 200300 or more years, trees growing at high

can be burned by intense fire, but rarely does fire kill

elevations are more likely to reproduce successfully from

all seeds of either type. Because the size of the can-

seeds dispersed into canopy gaps from nonserotinous

opy seed bank varies across the landscapeprimarily

cones. In contrast, at lower elevations, where typical fire

because of variation in the proportion of serotinous

intervals are less than 200 years, canopy gaps are less

trees and because of seed mortality during some fires

common and serotinous cones may be a more successful

the density of lodgepole pine seedlings after a fire varies

adaptation.

greatly. Seedling densities after the 1988 Yellowstone

As described previously, selection for serotinous

fires ranged from less than 5 per acre to greater than

and nonserotinous cones also may be influenced by

100,000 per acre. The fewest seedlings were found

seed predators, notably the red squirrel (see fig. 11.4).75

where the burned canopy was a mix of lodgepole pine

Considerably more seed is produced per year in non

and other species, essentially all lodgepole pine were

serotinous cones than in serotinous cones, a pattern

nonserotinous, and the fire burned through the canopy

that can be attributed to the need for allocating more

at high intensity. The most seedlings occurred where

energy to protective tissues in the serotinous cones

the burned canopy was essentially all lodgepole pine,

to minimize seed consumption by squirrels.76 Having

most trees were serotinous, and the fire scorched but did

both cone types in a stand seems adaptive for lodgepole

not consume all of the canopy.77

pine, as it ensures a seed source regardless of the type

As the seedlings grow into saplings and small trees,

of disturbance. Nonserotinous cones also enable the

the very dense post-fire stands become difficult to walk

establishment of lodgepole pine in nonforest habitats,

through. Such dense pine forests are commonly referred

Mountain Forests

to as doghair stands (fig. 11.16). The growth of individual


trees is slow in such places, but most survive and continue to produce seed for well beyond a century, though
remaining less than 4 inches in diameter. Not all lodgepole pine stands are so dense, however, as densities vary
greatly across the landscape. Historical reconstructions
in Yellowstone National Park revealed that, over time
spans of about 200 years, varying densities converge on
an average of about 500 trees per acre.78 Stands with different densities become more similar over time because
of self-thinning of initially dense stands and gradual
infilling of initially sparse stands (that is, by the establishment of young trees in older stands). Other disturbances affecting landscape patterns in lodgepole forests
include comandra blister rust, dwarf mistletoe, bark
beetles, root rot, and windstorms.
The effects of twentieth-century fire exclusion are
not nearly as significant in lodgepole pine forests as
in ponderosa pine and Douglas-fir forests, primarily because fires were infrequent and typically of
high severity in lodgepole pine forests even before
EuroAmericans ushered in fire suppression efforts.
Also, the fire regime is controlled more by weather and
climate than by ignition sources or variation in fuel
conditions.79 It is true that many fire ignitions have
been suppressed in the lodgepole pine zone, and that
some fires ignited in lower-elevation forests might
have burned up into the lodgepole had they not been
suppressed at the source. However, in most years such
fires would have been small, because the fuels at the
high elevations occupied by lodgepole pine historically were too wet for extensive fires. In dry years when
the weather is conducive to fire, the normal fuel struc-

Fig. 11.16.The trees in this dense stand of lodgepole pine


are small but are about 100 years oldprobably because the
previous forest that burned had a large number of trees with
serotinous cones. Also, the fire was not intense enough to
burn the cones and their seed, and the climatic conditions
after the fire were ideal for seedling establishment. With
about 6,000 trees per acre, there are few understory plants.
Unlike the dense ponderosa pine forest shown in fig. 11.8,
lodgepole pine forests such as this one are not caused by fire
suppression. Photo taken southwest of Dry Park in the Medicine Bow Mountains. Elevation 9,160 feet.

ture in lodgepole pine forests allows fires to spread


quickly and they are not readily suppressed. Had there
been no fire suppression during the twentieth century,

Spruce-Fir Forest

young lodgepole pine stands would be somewhat more

Engelmann spruce and subalpine fir often dominate

abundant today and old stands somewhat less sobut

forests at higher elevations and in cool ravines, most

todays distribution of stand ages does not differ greatly

likely because their seedlings tolerate the lower temper-

from what would have been seen in previous centuries.

atures and because more water is available longer into

Moreover, recent large fires in lodgepole pine forests

the summer than elsewhere (fig. 11.17).81 Both appear to

have created extensive stands of young forests (see

have lower water-use efficiencies than lodgepole, which

chapter 15). As discussed in the next chapter, climate

means that they require larger amounts of water for

change could increase the frequency and intensity of

every gram of new plant material they produce. Conse-

fires if the summers become drier, as predicted because

quently, the pine survives on drier sites at lower eleva-

of warmer temperatures.80

tions, where water stress is likely to develop. Ponderosa

193

194 Foothills and Mountains

Fires are less frequent in the spruce-fir zone than in


the lodgepole pine zone, because of even wetter conditions in the subalpine environment. Many spruce-fir
forests persist for centuries without fire. For example,
a tree-ring-based fire history study in a 75,000-acre
sub
alpine portion of Rocky Mountain National Park
recorded no significant fires during the previous 400
years in 27 percent of the forested area.82 This pattern
of few fires was true before the advent of efforts to put
them out, but those that did occur, and that occur
today, tend to be very hot and stand-replacing, just as
in lodgepole pine forests.
All four tree species in subalpine forestsspruce, fir,
aspen, and lodgepole pinemay become established
rapidly after a fire, depending on the availability of seed

Fig. 11.18.Occasionally subalpine fir and Engelmann spruce


develop clones, such as the two shown here. This occurs when
lower branches are pressed against the ground by snow and
new roots begin to grow, a process known as layering. A new
tree then develops from the rooted branch. Meadow plants at
this site in the Medicine Bow Mountains include alpine bistort, alpine timothy, and Idaho fescue. Elevation 10,400 feet.
Fig. 11.17. Subalpine forests dominated by Engelmann spruce
and subalpine fir are common at higher elevations, such as in
this area in the Medicine Bow Mountains. Note the presence
of downed wood, small and large trees, and the continuous
fuel ladder from the understory to the canopy. The spruce
beetle, a native insect, has recently killed many large trees in
this stand. The bright green plant in the understory is dwarf
huckleberry, an indicator of a relatively cool, moist environment. Elevation about 9,000 feet.

pine and Douglas-fir also are better adapted to the lowerelevation environment.
Unlike the pines and Douglas-fir, spruce and fir are
capable of vegetative reproduction when lower branches
are pressed to the ground by snow or a fallen log. The
branch then develops roots, known as adventitious
roots, after which the end of the branch begins to grow
upright into a new tree. Eventually the branchs connection to the parent tree is severed by death of the initiating branch and its decomposition. This process, known
as layering, produces clones of the original tree that are
commonly found locally at higher elevations. Clusters
of subalpine fir in meadows often can be attributed to
this cloning process (fig. 11.18).

Mountain Forests

and moisture.83 If lodgepole pine or aspen dominates the

spruce that the beetles generally do not kill. The growth

stand first, they will persist for a century or more. How-

of other plants is stimulated as well, for example, aspen,

ever, they are less tolerant of the understory environ

lodgepole pine, and numerous shrubs and forbs.

ment that typically develops at higher elevations and

Extensive areas of spruce-fir forest can also be affected

are gradually replaced by spruce and fir if no subsequent

by mortality of subalpine fir caused by the western bal-

stand-replacing disturbance occurs. One paleoecological

sam bark beetle (Dryocoetes) and its associated fungal

study, based on pollen analysis, found that dominance

pathogens. As with other bark beetles, warm, dry cli-

alternated over the centuries between lodgepole pine and

matic conditions favor this insect.88

spruce and firwhich may reflect post-fire dynamics or

The relative abundance of spruce and fir varies

shifting tree species distribution in response to climate

greatly from place to place. In stands older than about

changes.84 Spruce-fir forests also have been observed to

100 years, the largest and oldest trees typically are

develop on former meadows, sometimes following the

Engelmann spruce, as this species may live 500 years or

initial establishment of aspen.85 In much of the spruce-

longer.89 In contrast, subalpine fir is usually more abun-

fir zone, however, lodgepole pine and aspen are not pres-

dant, but the trees are smaller and younger, rarely more

ent. In such places, spruce and fir become re-established

than 250 years old. Fir seedlings and saplings may also

in the burned area without any competition from aspen

be 1020 times more numerous than those of spruce.

or lodgepole pine. Research in northern Colorado sug-

Apparently the roots of fir seedlings are better able to

gests that after 100200 years, on some sites, spruce no

penetrate the considerable litter that accumulates on the

longer reproduces in the forest understorya period

forest floor, whereas spruce seedlings are usually found

described as the spruce exclusion phase. After another

where mineral soil has been exposed, such as around

100 years, many of the dominant fir and spruce begin

the tipped root system of fallen trees. Some investiga-

to die, creating canopy gaps that initiate the spruce re-

tors have also observed spruce seedlings on rotting logs

initiation phase. The final phase is a second-generation

at sites where the decomposing wood remains wet well

spruce-fir forest which persists until the next severe dis-

into the summer, such as in moist ravines.90 Spruce pro-

turbance.

duces especially large numbers of seed every 25 years,

86

The likelihood of such a long interval with

no major disturbances caused by blowdown or insect


outbreak is low.

fir about every 3 years.


That subalpine fir reproduces more effectively than

Insect outbreaks also shape the structure and dy

spruce in the forest environment raises the question

namics of spruce-fir forests. Especially important is

of how spruce persists as a co-dominant. The answer

the spruce beetle, a bark beetle with a life cycle simi-

lies in their different life history strategies.91 Fir is

lar to the mountain pine beetle (described earlier in

shorter lived but produces many seedlings, whereas

this chapter).87 The spruce beetle attacks older stands

Engelmann spruce compensates for generally poor

with a high proportion of Engelmann spruce in the

reproductive success with increased longevity. Thus,

overstory. Susceptible stands usually display slow tree

the greater proportion of subalpine fir seedlings and

growth, suggesting that the trees are of low vigor and

saplings does not mean that fir will eventually become

are unable to produce sufficient resin to exclude the

the sole dominant.

beetles. Unlike mountain pine beetles, which reproduce

Engelmann spruce and subalpine fir are both com-

successfully only in live trees, spruce beetles can pro-

paratively tolerant of the understory environment; that

duce large numbers of offspring in trees that have died

is, young trees are tolerant of shade and competition for

recently from other causes. Moreover, an abundance of

water and nutrients. Therefore, the two species usually

dead trees, whether from logging or windthrow, may

develop uneven-aged forests, with both young and old

provide the energy base for the initial development of

trees. Because the overstory trees are the same species

a spruce beetle outbreak that spreads subsequently into

as in the understory, it is tempting to view spruce-fir

live trees. The infestation opens the canopy, greatly

forests as being in a state of equilibrium, with dying

accelerating the growth of the abundant subalpine fir

overstory trees being replaced with younger trees of the

(which is not a host for the spruce beetle) and the small

same species. However, all forests are subject to ongoing

195

196 Foothills and Mountains

changes caused by fire, insects, diseases, windthrow,


and other factors.
As with lodgepole pine forests, twentieth-century
fire exclusion had a relatively minor influence on the
structure and dynamics of spruce-fir forests. Timber
harvesting has been less widespread in subalpine forests; consequently, they appear today much as they
would have in the early 1800s.

Aspen Forest
Aspen forests extend from the foothills to the subalpine
zone, typically in depressions, ravines, valley bottoms,
and the lee sides of ridgeswhere water availability is a
little higher. Because of this distribution pattern, many
ecologists have assumed that aspen requires more water
than other trees do. In fact, adult aspen use considerably less water per unit leaf area than lodgepole pine,
subalpine fir, and Engelmann spruce.92 The general
restriction of aspen to wetter areas is probably due more
to the intolerance of aspen seedlings to drought than to
the intolerance of mature trees.
Aspen is uniqueit is the only deciduous tree in the
region that grows in an upland environment that seems
to favor evergreen plants. The loss of all leaves each fall
is not efficient with regard to nutrient conservation, and
this may partially explain why aspen seems restricted to
depressions, where nutrients are more readily available.
As with conifers, aspen twigs are probably capable of
reabsorbing nutrients from the leaves before they fall,

Fig. 11.19. Some aspen forests have an understory of suba lpine


fir or, in this area on the west slope of Wyomings Sierra
Madre, lodgepole pine. In the absence of another fire or harvesting of the conifers, the area dominated by aspen gradually
becomes smaller and conifer-dominated forests become more
extensive. Photo taken after leaf fall. Elevation 8,500 feet.

an important mechanism for retaining at least some


of the limiting nutrients in the leaves. Also, although

In Wyoming, aspen usually occurs in small groves

aspen loses its leaves in the fall, there is chlorophyll in

(figs. 11.19 and 11.20). However, extensive forests are

the barkeasily seen after scratching the bark. It was

found in a few areas, notably on the west slope of the

previously thought that bark photosynthesis was an

Sierra Madre south of Rawlins.94 Aspen abundance there

important component of the trees total annual pho-

may be enabled by the Arizona Monsoon, which brings

tosynthesis, especially during the leafless period. How-

more reliable summer precipitation to this part of the

ever, recent research indicates that little carbon dioxide

state (see chapter 3). Southward into Colorado and New

penetrates the transparent but corky bark outside the

Mexico, aspen becomes increasingly common, perhaps

layer of chlorophyll. Instead, the chlorophyll in the

because of more summer and fall rain. There, it tends

bark converts some of the carbon dioxide produced by

to occupy the same topographic positions as lodgepole

the trees own internal respirationa process known

pine does in the Central and Northern Rockies.95

as refixation. In this way, the bark contributes to the

The tendency for aspen to occur in small groves in

trees energy budget.93 The plant is effectively evergreen

Wyoming can be explained in part by its clonal nature,

despite having deciduous leaves. As noted previously,

with new trees developing from root sprouts, also

dwarf huckleberry also is a deciduous evergreen.

known as suckers. Seedling establishment is rare, not

Mountain Forests
Fig. 11.20. Aspen groves in
Wyoming tend to be small,
except on the southwestern
slopes of the Sierra Madre,
east of Baggs. Here the aspen
grow over large areas, similar
to the aspen farther south in
Colorado and New Mexico.
Patches of Gambel oak also
are found in this area. Note
that some of the aspen clones
are still green (see text for
explanation).

because there is a lack of seed but because the environ-

the ramets of some clones (genets) may intermingle.97

96

mental requirements for establishment are rarely met.

Genetic diversity undoubtedly explains why aspen is

Aspen seedlings are delicate and cannot tolerate even a

found over such wide environmental gradients in the

brief period of desiccation. They often die before their

Rocky Mountain region and, in fact, why it is the most

roots can penetrate through forest floor litter to the

widespread tree in North America.

mineral soil. Also, they may be unable to compete with

Because of aspens ability to sprout, it dominates

other plants for limited water. Whatever the mecha-

temporarily after severe disturbances in many conifer-

nisms that limit seedling establishment in undisturbed

ous forests, most likely because the sprouts have more

forests, aspen seedlings have been observed in recently

stored energy available for rapid growth than do conifer

burned forests where mineral soil is exposed and poten-

seedlings. More than 10,000 sprouts per acre may reach

tially competing plants have been temporarily reduced,

a height of 3 feet or more within a year following a fire.

such as in Yellowstone National Park after the 1988 fires

Conifers, especially subalpine fir, sometimes persist

(see chapter 15). Seedlings also have appeared on sedi-

in the understory and eventually overtop the aspen

ments exposed in abandoned beaver ponds.

because its seedlings become established on the aspen

With sprouts as the principal form of regeneration,

forest floor more readily than do those of other trees

nearly all groves develop as a genetically uniform clone

(see fig. 11.19). Lodgepole pine and Engelmann spruce

or a mix of several discrete clones. Essentially, what

also replace aspen under some conditions.98

nearly everyone refers to as a tree is more appropri-

Where other trees do not become established, aspen

ately termed a branch or shoot. Botanists use the term

continues to dominate stands indefinitely, with new

ramet. Although ramets typically die after about 100

sprouts replacing the senescing older stems. Generally,

years, the plant itselfthe genetlives for much longer,

most aspen sprouts do not survive in the understory

most likely for centuries. Thus, more genetic variabil-

shade of aspenan interesting anomaly, as it would

ity exists between aspen groves than within one grove,

seem that the small trees could tap the energy stored

as is apparent when, for example, two adjacent clones

by the larger trees (which have leaves fully exposed to

lose their leaves at different times in the fall. Of course,

sunlight). Plant hormones produced by the dominant

197

198 Foothills and Mountains

shoots apparently suppress the growth of the smaller

of the conifers, more prescribed fires, or reductions in

ones belowa case of apical dominance. Aspen stands

ungulate numbers.

burn periodically, at intervals of decades or centuries,


and the thin-barked adult stems are usually killed.99
However, abundant post-fire sprouting and rapid growth

Mountain Forests and Wildlife

result in the re-establishment of an aspen grove within

The forest-meadow-wetland mosaic provides essen-

a few decades. Chronic heavy browsing on the aspen

tial habitat for numerous birds and mammals. A few

sprouts by wild and domestic ungulates can inhibit this

are found there all year long, such as the blue grouse,

process of aspen regeneration (see chapters 10 and 15).100

boreal owl, Clarks nutcracker, gray jay, red crossbill,

Old landscape photos from the late 1800s and early

beaver, lynx, marten, red squirrel, snowshoe hare, and

1900s commonly show more aspen than is visible in the

voles. Others migrate to the foothills for the winter,

same scenes today, suggesting that aspen is declining

escaping deep snow and colder temperatures, returning

in at least some areas. This comparison may be mis-

to the mountains in the summer for more reliable for-

leading, simply because the late 1800s was a time of

age and better hiding cover. Elk, mule deer, and cou-

extensive burning throughout the West, and post-fire

gar are notable examples. Many birds, such as warblers,

aspen probably was exceptionally abundant around

thrushes, kinglets, and white-crowned sparrow, migrate

1900, compared with earlier in the nineteenth century

to warmer southern climates for the winter, along with

or in previous centuries. Thus, aspen abundance at the

the few ducks and mergansers that use mountain lakes

beginning of the twentieth century is not necessarily

and beaver ponds. As noted, some birds and small mam-

an appropriate historical benchmark for land manage-

mals rely on conifer seed for their food. Others feed on

ment.101 Moreover, photos from other areas show that

mushrooms or mature insects, or dig out insect larvae

aspen cover remained stable or even increased in some

from decomposing wood or the bark of living trees. The

places during the twentieth century.

Some resource

abundance of insects emerging from wetlands provides

managers are concerned that twentieth-century fire

food for birds that are agile enough to catch them. Even

exclusion and other factors have led to deterioration of

bears are known to eat moth larvae, which can be abun-

many aspen groves, as the adult aspen stems gradually

dant under the rocks on talus slopes.

102

die and are replaced by conifers or shrubs instead of

Commonly, mountain wildlife is secretive and not

new aspen sprouts. Aspen shoots can be killed by vari-

easily observed. As in other ecosystems, the best time

ous pathogens, including root rot and canker diseases,

to find them is in the spring, when the birds can be

especially if the bark is damaged. This trend of stand

located by their singing, or during the winter, when

deterioration is especially worrisome if the capacity for

many species congregate in flocks, or footprints can be

sprouting diminishes as stands age, or if the intensity of

observed in the snow. Aspen groves are often identified

ungulate browsing or beaver cutting is high (see chap-

as hotspots of biological diversity, perhaps because of

ters 10 and 15).103

the abundance and diversity of understory plants, the

The first decade of the twenty-first century saw exten-

frequent proximity of aspen to water, and the relative

sive death of aspen trees in Wyoming and across much

ease with which cavities can be found or created for

of the West, apparently triggered by the severe drought

nesting. In all kinds of forest, many animals depend on

of 20002003.104 The patterns and implications of this

dead trees, whether for perches or cavities.106 For many

mortality event, referred to as sudden aspen decline (SAD),

species, the juxtaposition of wetlands, meadows of vari-

are discussed in chapter 12. Despite such developments

ous kinds, and forests of different ages and types makes

in aspen forests, there seems little need to be concerned

for the perfect habitat. As always, the landscape mosaic

about aspen extinction.105 Still, the tree could become

must be considered, not just a few acres of meadow or

less abundant in some areas as aspen-dominated forests

forest.

gradually change to conifer-dominated forests or shrub-

Overall, forests of all kinds cover about 15 percent of

lands. To maintain the current abundance of aspen


forests in the landscape may require more harvesting

Wyoming and are found primarily in the mountains,


where temperature, moisture, and nutrient conditions

Mountain Forests

are sufficiently favorable to enable tree seedling estab-

amenities have led to intensive management in some

lishment and growth. Although only six tree species are

areas, and to concerns about whether current manage-

common in these forests, their adaptations vary consid-

ment practices will sustain wildlife populations as well

erably, and the influences of fires, insects, windstorms,

as wood production. Recent large fires, insect outbreaks,

and abrupt topographic changes create a mosaic that

and the specter of climate change raise questions about

includes a diversity of forest types intermingled with

the resilience of Rocky Mountain forest ecosystems, at

meadows and shrublands, along with mountain lakes

least as we know them todaya topic that is discussed

and alpine tundra. Demands for forest products and

further in the next chapter.

199

Chapter 12

The Forest Ecosystem

Thinking about forest ecosystems is less intuitive than

thousands of gallons of water and tons of chemicals

considering the adaptations and relative abundance of

streaming upward through tree trunks, photons of

different kinds of plants and animals, as in the previous

energy absorbed by leaves and put to work evaporat-

chapter. However, ecosystem science enables a better

ing water through leaves and fixing energy in organic

understanding of the factors determining tree growth,

compounds, food manufactured in leaves streaming

the development and maintenance of soils that allow

to growing points, insect predators quietly nibbling

that growth, the amount and quality of water flowing

away, rocks being broken down into useable nutri-

from watersheds, the effects of disturbances (such as

ents, microbes disassembling organic compounds

fires, insect outbreaks, and timber harvesting), and the

and freeing nutrients for reuse, all species playing

effects of climate change. Curiously, dead plants and

out their roles in reproduction, the forest ecosystem

animals are just as important as those still alive, if only

grudgingly restocking the forest stream with water,

because they provide microhabitats for various forms of

and a million other things.1

life. In some cases the dead plant materialthe O horizon of the soil (commonly referred to as the forest floor,
litter, or debris)functions as an important mechanism
for erosion control and for conserving nutrients that
may be limiting.
The intrigue of forest ecosystems, or any ecosystem
for that matter, is that the important processes are not
visible. To obtain data, scientists use various instruments to measure processes such as solar radiation (in
the wavelengths used for photosynthesis), the rates of
plant growth and litter decomposition, and the movement of nutrients and water. Ones perspective on a
forest changes when such information is revealed. The
late F. Herbert Bormann, eminent ecosystem scientist at
Yale University, wrote:

Energy Flow, Productivity, and Carbon Sequestration


Most of the suns energy reaching a forest, or any other
terrestrial ecosystem, is used for evaporating water and
heating the environment; a mere 13 percent is used for
photosynthesis. In mountain forests, because of high
elevation and a comparatively thin atmosphere with
low humidity, much of the heat gained during the day
is lost at night by radiation back into space. The result is
a relatively cool environment. Also, water is a limiting
factor in many places by mid- to late summer. Consequently, because of a cool, relatively dry environment,
tree growth rates are low in the Rocky Mountains compared to rates in other places in North America, such

Today when I visit a forest, there is . . . a sense of

as the Pacific Northwest and the Southeastern Coastal

being surrounded by an enormous dynamism:

Plain (fig. 12.1). Annual variation in wood produc-

200

The Forest Ecosystem 201

NPP (t/ha/yr)

Live tree
biomass
15
Scots pine

10

Atmospheric
deposition
0.3

Slash pine
5

Lodgepole pine

Fixation
0

25

50

0.01

0.6

Aboveground
litter

1.2

75

STAND AGE (years)


Fig. 12.1. Rate of forest plant growth, mostly tree growth
but also that of understory shrubs and herbaceous plants, is
known as net primary productivity (NPP)expressed here as
metric tons per hectare per year. NPP changes significantly as
forests age. Note the low growth rate for Wyoming lodgepole pine forests compared to slash pine in Florida and Scots
pine in northern Europe. One metric ton = 2,205 pounds;
1 hectare = 2.47 acres. Adapted from Knight (1991).

0.9

Fixation

0.01

Soil
organic
matter

tion can be observed in the annual rings of trees and


is caused mostly by annual variation in growing season temperature and precipitation. Also, younger forests tend to have higher growth rates than older forests.
Growth rate, also known as net primary productivity
(often abbreviated as NPP), is commonly expressed as
the average rate at which all new plant biomass is added
to the forest annually, whether in the form of trees,
shrubs, grasses, or forbs.2
Living trees, soil organic matter, and dead plant material on the forest floor are the largest components of
the forest biomass and change through time after a dis
turbance (figs. 12.2 and 12.3). The forest floor biomass is
composed almost entirely of leaves, twigs, branches, and
the boles of fallen trees. Some of itnot allis burned
during a forest fire, but the litter accumulates in total
weight for many years. This litter buildup occurs because
winters are long; summers are relatively dry; and most
plant tissues in Rocky Mountain environments have high
lignin content and high carbon-to-nitrogen ratios, both
of which slow the rate at which bacteria and fungithe
primary decomposerscan convert organic materials
to inorganic molecules. Microbial decomposition is also
limited by the relatively cool and sometimes dry conditions that prevail in the summer. Notably, considerable

Outflow
0.01
Fig. 12.2.Carbon and nitrogen distribution in a lodgepole
pine forest. The size of the boxes indicates the relative
amounts of organic matter in live tree biomass, aboveground
litter, and soil organic matter (which includes bacteria and
fungal biomass but not live root biomass). Roots are part
of the live tree biomass box. Shrub and herbaceous plant
biomass is not shown but is less than 1 percent of the tree
biomass. Annual nitrogen flows are shown by the arrows;
the numbers are grams per square meter per year. Much of
the nitrogen in the rooting zone is in organic matter that
is not readily decomposable. Note that nitrogen inputs are
larger than nitrogen losses, suggesting that nitrogen is accumulating in the tree biomass, litter, and/or soil organic
matter, probably because it is a limiting factor for plants and
microbial organisms. Also, note that the tree uptake estimate
is larger than the sum of the input estimates to the rooting
zone, which suggests that the amount of soil nitrogen is
gradually depleted in this forest as biomass accumulates. The
soil nitrogen pool is replenished as the forest ages further or
following such disturbances as fire. Based on data in Fahey et
al. (1985) and Fahey and Knight (1986).

202 Foothills and Mountains

Biomass Increment (t/ha/yr)

(such as the red squirrel and red-backed vole) commonly

Live trees

Forest floor litter


Dead trees and
downed wood

graze on mushrooms, truffles, and puffballs, the reproductive structures of fungi that decompose the detritus,
and birds commonly prey on these small mammals and
various invertebrate detritivores. Moreover, evidence
suggests that these animals are important in dispersing

the spores of fungi, which in turn are important not


only for decomposition butin the case of mycorrhizal

fungialso for the establishment and growth of new


plant seedlings, including those of all Rocky Mountain
trees.4

-1

The importance of mycorrhizal fungi is suggested


by estimates that up to 15 percent of the net primary

-2

40

80

120

Stand Age (years)

160

200

Fig. 12.3. Rates of biomass increment or decline in live trees,


forest floor litter, and dead trees and downed wood change
as forests grow to maturity and become old-growth forests.
Older forests have more dead trees and downed wood, but
tree growth slows as forests age (see also fig. 12.1). Dead trees
and downed wood, sometimes referred to as snags and coarse
woody debris, decline soon after fire, as decomposition rates
of this material are relatively high when water and nutrients
are more readily available for the decomposers. The forest
floor litter increases initially after a firewhich commonly
burns much of the pre-fire forest floorbut then remains
about the same, as litterfall becomes approximately equal to
the rate of litter decay. The patterns illustrated will be somewhat different for forests growing on different soils, at different elevations, or in different topographic positions. These
graphs are based on data collected from several lodgepole pine
forests at an elevation of about 9,500 feet in the Medicine Bow
Mountains. Rates of change are expressed as metric tons per
hectare per year; see fig. 12.1 for unit conversions. Adapted
from Pearson et al. (1987); see also Smith and Resh (1999) and
Kashian et al. (2013).

productivity in a coniferous forest goes to the maintenance of these fungi on roots.5 All studies thus far
indicate that, rather than being parasites, mycorrhizal
fungi develop a mutually beneficial association with
their hosts.6 The fungi derive energy from the plants in
the form of carbohydrates, while the fungal filaments,
known as hyphae, extend beyond the roots and enhance
water and nutrient uptake. The decaying biomass of the
fine roots and associated fungi may contribute more to
nutrient availability than do decaying leaves, twigs, and
branches.
On average, only 2 percent or less of the energy fixed
by plants during photosynthesis flows through animals
in terrestrial ecosystems, whether forests, shrublands, or
grasslands, yet animals often influence the ecosystem
in important ways. An example is the planting of whitebark and limber pine seeds by Clarks nutcracker (see
chapters 10 and 11). The populations of different herbivores and carnivores continually fluctuate, with the
result that their influences are greatly amplified when
they are abundant, such as during outbreaks of bark

decomposition occurs under snow during the winter, as

beetles. A less well-known example occurs when the

some of the decomposers are able to function at the soil/

populations of porcupines are high. Porcupines feed on

snow interface.3 Because detritus accumulates over time

tree bark, resulting in reduced flow of carbohydrates to

and is flammable when dry, fire is another important

the roots, which lowers plant growth in the forest as a

pathway for energy flowa decomposition process that,

whole.7 Ecosystems are dynamic, with all components

like bacteria and fungi, releases the inorganic nutrients

changing over time, as does the amount of energy flow-

required for new plant growth (fig. 12.4).

ing through the detrital, grazing, and fire pathways.

The combined biomass of forest floor litter, fungi,


and bacteria is far greater than the total biomass of
vertebrates and invertebrates. Not surprisingly, various

Hydrology of Forest Landscapes

animals have evolved to depend on detritus for a signifi-

Mountain landscapes are hydrologically distinctive

cant part of their energy. For example, small mammals

from the surrounding lowlands, because they receive

The Forest Ecosystem 203

Fig. 12.4.Importance of nutrients and water for the growth


of lodgepole pine. In the photo on the left, an intense fire
has burned the forest floor and some of the downed wood
buried in the litter. The photo shows a pine root that had
grown horizontally in the top layers of the soil. When the
root encountered a decomposing log, now consumed by the
fire, it turned 90 degrees as it grew into the logmost likely
because the log provided a better source of nutrients or water.
In this case, the log was burned but not the root. In the photo
on the right, a dense stand of new pines is developing after
a clearcut that occurred about 12 years previously. However,

adjacent to the uncut forest is a band with very little vegetation around the entire perimeter of the clearcut, even with
full sunlight. This pattern is common and suggests that the
roots of the mature forest trees grow outward from the trunk,
to a distance of about 12 feet, and that the roots provide too
much competition for the establishment of tree seedlings and
other plants. Competition for water and nutrients may be
more important than competition for light in the understory
of lodgepole pine forests. The trees in the uncut forest on the
left are small because of their high density but are about 100
years old (see fig. 11.16).

more water than can evaporate during the year. In other

barely sufficient to wet the forest floor, to say nothing

words, the ratio between precipitation and evaporation,

of wetting the soil. In contrast, snowmelt normally

known as the P-E ratio, is greater than one. Conse-

saturates the soil, and then soil moisture declines

quently, the mountains are the regions primary source

more or less steadily during the summer and early fall,

of river water and groundwater, on which agriculture,

whether the vegetation is a forest, woodland, meadow,

industry, and municipalities depend as well as ripar-

or alpine tundra. 8 The actual amount of surface water

ian and wetland ecosystems. The mountains have been

draining from a watershed depends on at least seven

referred to as the water towers of the West.

factors: water content of the snowpack, potential for

Because snow historically has accounted for 5075

snow drifting, type of vegetation and the amount of

percent of the annual precipitation in the mountains,

water transpired, water-holding capacity of the soil,

streamflow has a pronounced flood peak during snow-

weather conditions, proportion of water that perco-

melt in the spring (fig. 12.5). Summer rains normally

lates into groundwater, and patchiness of the land-

contribute little to streamflow, because most rains are

scape mosaic.

STREAMFLOW (cfs)

204 Foothills and Mountains

In contrast, forests produce larger amounts of leaf

8
7
6
5
4
3
2
1

area and have higher rates of transpiration. Figure 12.6

1956-1971

compares the hydrologic budget of a typical lodgepole


pine forest in the Medicine Bow Mountains with that
of an old-growth Douglas-fir forest in Oregon (which
receives far more annual precipitation, though less

1940-1955
APR

MAY

JUNE

snow). In the Oregon forest, about 35 percent of the


JULY

AUG

SEPT

Fig. 12.5.Typical hydrograph (solid line) for a Rocky Mountain creek, showing how streamflow increases dramatically in
the spring when the snow is melting. More streamflow commonly occurs following timber harvesting in small patches
(dotted line). The units are cubic feet per second (cfs). Adapted
from Alexander (1987a).

annual precipitation evaporates and transpires during


a typical year, and 65 percent flows to the stream or
groundwater. Evaporation occurs primarily because
of interception. In both regions, interception varies
according to the amount of plant and litter surface area
exposed abovegroundwhether rainfall occurs as a
slow drizzle or quick thunderstormand wind velocity at the time of snowfall.10 In the Rocky Mountains,

Snow
The amount of water in snow cannot be predicted from
snow depth because snow can be comparatively wet or
dry. For example, a snowpack that is 3 feet deep may
have the equivalent of as little as 6 inches or as much
as 16 inches of liquid water.9 Snow drifting is typical,
and thus there can be considerable spatial variability in
depth as well as water content. Because of the importance of snow as a source of water for irrigation and
downstream reservoirs, the Natural Resources Conservation Service and other agencies routinely monitor the
water content of mountain snow. By combining such
data with information about the watershed, the amount
of runoff that will be available downstream during the
summer can be estimated.

Vegetation
Two characteristics of Rocky Mountain vegetation that

sublimation from snow may be more important than


interception.
Leaf area and other stand characteristics vary considerably from one forest to another.11 Not surprisingly,
the amount of leaf area tends to increase with water and
nutrient availability at specific locations.12 In a range of
lodgepole pine forest conditions in the Medicine Bow
Mountains, actual evapotranspiration for the period
from early spring to late fallwhen probably 95 percent
or more of evapotranspiration occursvaried from 10
to 20 inches.13 This represents 3395 percent (a mean of
75 percent) of the total annual precipitation. Outflow to
streams and groundwater accounted for 080 percent of
the snow water, depending on weather conditions during
the year and on forest characteristics, such as the leaf area
index and soil storage capacity. Almost no outflow occurs
during years of relatively low snow-water equivalent in
forest stands with substantial amounts of evergreen leaf
area capable of transpiration in the early spring.

strongly affect runoff, whether that of a meadow or a for-

The importance of leaf area is further illustrated by

est, are the total amount of leaf area and leaf type (ever-

another study done in the Medicine Bow Mountains,

green or deciduous). Leaves are important because water

where the water budgets of two adjacent but differ-

vapor is lost through their stomata (a process known

ent lodgepole pine forests were determined. One was

as transpiration), and also because water adheres to the

a dense doghair stand with about 6,000 trees per acre,

exterior of the leaves and evaporates without becoming

and the other a more open stand with about 800 trees

available to the plants (a process known as interception).

per acre. Both originated after the same fire in the late

The effect of vegetation on water outflow is least when

1800s. The two differed in tree density and biomass, but

the leaf area is low and the leaves are deciduous, such as

both stands had the same leaf area and about the same

in a meadow; probably 8090 percent of the snow water

transpiration rates.14

accumulating in meadows becomes part of streamflow


or groundwater.

Significantly, there are differences between the


hydrology of lodgepole pine forests and spruce-fir for-

The Forest Ecosystem 205

Oregon
Douglas-fir Forest
Precipitation
237

Transpiration

Evaporation

Evaporation
5
Snow
2
43

218
37
Canopy
storage

Wyoming
Lodgepole Pine Forest

24

198

4
Canopy
storage

4
Litter

25

18

38

Rooting
zone
13

155

Outflow

Outflow
Fig. 12.6. Annual hydrologic budgets for stands of raindominated Douglas-fir in Oregon and drier, snow-dominated
lodgepole pine in Wyoming. The boxes represent major storage compartments; the arrows are processes affecting water
movement through the ecosystem. Wide arrows indicate
processes that affect relatively more water than processes
with narrow arrows. Numbers are centimeters of water (for
example, 25 cm is the volume of water required to cover any

22

40

Snow

181

Rooting
zone

Evaporation

Evaporation and
Sublimation

14
Litter

Precipitation
66

Transpiration

area 25 cm deep; to convert centimeters to inches, divide by


2.54). Canopy interception is higher in the Oregon forest,
because more rainfall and a higher leaf area index occur there
compared to Wyoming forests. Interception and subsequent
evaporation of summer rains is important in Wyoming, as is
sublimation of snow during winter. The figure for Douglas-fir
is adapted from Sollins et al. (1980); data for lodgepole pine
are from Knight et al. (1985) and Biederman et al. (2012).

ests.15 Forests dominated by spruce and fir in the Rocky

coniferous forests during long periods without distur-

Mountains have considerably more leaf area than do

bance (see chapter 11), with streamflow decreasing as

lodgepole pine forests, and furthermore, the spruce

this shift occurs.16

and fir use water at higher rates per unit of leaf area

Reductions in leaf area stemming from tree cutting,

than either pine or aspen. All things equal, streamflow

fire, and some other disturbances reduce evapotranspi-

would be less from watersheds where spruce-fir forests

ration and, therefore, potentially increase streamflow.17

are common.

Streamflow may continue at an elevated level for up to

Aspen forests are different because the leaves are

60 years, depending on forest type, intensity of distur-

deciduous and absent during most of the snowmelt

bance, and rate of vegetation regrowth, before declin-

period, precluding most early-spring transpiration.

ing to its predisturbance volume after the original leaf

Moreover, the leaves are shed in the fall, eliminating

area has been restored.18 Notably, for reasons described

the possibility of foliar transpiration during subsequent

below, the tree mortality caused by bark beetles does

warm days. Also, there is less potential for intercept-

not necessarily lead to an increase in streamflow.

ing snow on leaf surfaces, from which evaporation


and sublimation can occur. Some land managers have
noticed that the flow of mountain streams can increase

Soil Water-Holding Capacity

after aspen leaf fall. The net effect of forest domina-

Another variable affecting the amount of surface runoff

tion by deciduous trees is that less water is transpired

is the amount of water that the soil can store, that is,

and a larger portion of annual precipitation is available

the amount that is not lost to downward flow caused

for streamflow. Aspen forests may change to evergreen

by gravity. At some point, the soil becomes saturated

206 Foothills and Mountains

during most years, and additional water leaves the eco-

of less direct sunlight. In contrast, the snow on south-

system as surface or subsurface runoff. Thus, soil storage

facing slopes, especially at lower elevations, often melts

capacity at any particular time and place depends on

between snowstorms, with a considerable amount of

the amount of water already in the soil, with deeper and

the soil moisture evaporating on relatively warm days

finer-textured soils holding considerably more water

before additional snowfall occurs. Such periodic melt-

than shallow, coarse-textured soils. The uptake of water

ing makes the snowfall events analogous to summer

by plants leads to soil drying, enabling storage of addi-

showers, with each event producing too little moisture

tional water when it becomes available. The amount of

to percolate through the litter and rooting zone to cause

available storage generally increases during the sum-

substantial runoff. Thus, north-facing slopes contribute

mer, a function of the drying rate, rooting depth, soil

more to streamflow per unit area than do south slopes.

depth and texture, and the time since the last significant wetting eventusually snowmelt. By the end of
summer, the soil moisture of lodgepole pine forests may
be depleted to depths of 6 feet or more, as the pines have
taproots extending that deep.19

Percolation to Groundwater
Surface runoff in the form of streamflow is greatly
affected by the geologic substrate and by whether percolation into aquifers is possible. Granites and other igneous and metamorphic rocks are usually impervious to

Weather and Climate


Runoff from mountain watersheds is also affected by
weather. Gradual warming in the spring causes a prolonged snowmelt, which provides additional time
for early-spring transpiration, thereby reducing the
amount of water available for drainage to streamflow or
groundwater. In contrast, some springs are cold, becoming warm only when day lengths are long. Under such
conditions, the snow melts late and rapidly, and there is

water penetration, which enables more runoff. In contrast, watersheds underlain by sedimentary rock may
contribute significantly to the groundwater of intermountain basins, which is frequently tapped by wells.
Plants growing at the juncture of impervious bedrock
above and permeable sedimentary rock below benefit
from the higher amount of runoff (see fig. 10.3).

Patchiness of Landscape Mosaic

little time for spring transpiration. Moreover, the rate of

Heterogeneity or patchiness in landscapes, caused by

snowmelt occasionally exceeds the maximum infiltra-

disturbances and abrupt changes in various environ-

tion rate, causing overland flow and a higher potential

mental factors, has significant effects on watershed

for flooding, especially if the soil is already saturated

hydrology. For example, a landscape where forests are

or has become compactedas can happen when heavy

frequently interrupted by small meadows contributes

machines are used for timber harvesting. Water flowing

more water to streamflow than does an area of continu-

overland is not available for uptake by plants and can be

ous forest, primarily because wind causes snow drifting

a significant cause of soil erosion.

in such openings, where there is little leaf area and less

Fall weather may be as important as that of the

evapotranspiration.20 Furthermore, with more direct

spring. For example, heavy fall rains can saturate the

sunlight, snowmelt is more rapid in meadows, reduc-

soil just before the onset of winter, when evapotrans-

ing the time available for evapotranspiration. Con-

piration rates are low. Little storage capacity then exists

sequently, more snow water flows down beyond the

during the following spring, leading to the potential for

rooting zone than in areas uniformly covered by forest.

greater spring runoff. Also, if snow falls before the soil

Watershed managers sometimes recommend numer-

is frozen (and the snow is deep enough), the mountain

ous small patch cuts to increase the water yield from a

soils remain unfrozen for the entire winter and infiltra-

watershed. Based on research in northern Colorado, the

tion is enhanced.

largest increases in water yield are reached when 3040

Another factor causing variation in outflow is micro-

percent of a watershed is harvested in patches no larger

climate. For example, north-facing slopes in the North-

than 25 acres. Larger openings can have a different

ern Hemisphere are more likely to retain snow because

effect, because the stronger winds in the opening blow

The Forest Ecosystem 207

more snow into the downwind forest, where the snow

potentially altering the chemical composition of water

melts more slowly and higher evapotranspiration may

downstream.

occur.

Inputs or additions of nutrients occur as the result of

In general, abrupt changes between forests of dif-

dryfall (dust), wetfall (rain and snow), rock weathering,

ferent ages, or between forests and other kinds of veg-

animal immigration, surface or subsurface run-on, and

etation (such as meadows), have significant effects on

nitrogen fixation (see fig. 12.2). Losses occur with soil

snow distribution and melting. Weather and vegeta-

erosion, leaching, animal emigration, and, for nitro-

tion are all subject to rapid change (see chapter 11), and

gen, loss to the atmosphere. The conversion of nitrate

these changes affect the amount of streamflow, evapo

back into atmospheric nitrogen is a microbial process,

transpiration, and sublimation.

known as denitrification. Research in lodgepole pine


forests reveals that nitrogen and sometimes potassium

Nutrient Cycling in Forest Landscapes

and phosphorus accumulate in undisturbed forests over


time, even during years of heavy snowpack and large

Unlike water and energy, nutrients cycle in an area for

volumes of water outflow.22 Uptake by plants probably

prolonged periods. Some are lost from the ecosystem,

is especially effective for nutrient retention when the

but they can also be replaced. The nutrient that has

plants are growing rapidly, incorporating additional

received the greatest attention in Rocky Mountain conif-

nutrients into new tissues. 23

erous forests is nitrogen, primarily because this element


often limits rates of plant growth and decomposition,

Dead wood is a common component of almost all


forests and plays a key role in many ecosystem processes,

Nitrogen is the most abundant element in the atmo-

including nutrient cycling and carbon sequestration. As

sphere (78 percent), where it occurs as N2 a form that

fig. 12.7 shows, the nitrogen content of freshly decay-

normally must be converted to inorganic ammonium

ing wood increases considerably over 5 years or more,

or nitrate before it can be taken up by plant roots. Cer-

probably because of nitrogen uptake by the fungi and

tain microbes living in the root nodules of some plant

bacteria that are using the carbohydrates in the wood

species, such as lupine, or microbes that are free-living

as their source of energy. The additional nitrogen most

in the soil, are capable of making this conversion

likely comes from two sources, namely, atmospheric

known as nitrogen fixation. Some fixation also occurs

deposition and transport by fungal filaments connect-

with lightning. Ammonium and nitrate are released

ing the log to the soil. Subsequently, nitrogen content in

during the process of decomposition as well. Notably,

the wood declines as decomposition continues, which

some plants take up nitrogen from the soil in organic


agricultural and industrial technologies have greatly
increased the amount of nitrogen available to eco
systems, which is considered a significant environmental problem on Earth as a whole (see chapter 14).
The potential for nutrient leaching from the soil is
of interest in mountain landscapes, because so much
of the water washes through the soil during the 3- to
6-week snowmelt period. That nutrients are retained
during this spring flush is remarkable, considering that
a large portion of litter decomposition occurs during
the winter just before the spring thaw. Essentially,
nitrogen-limited plants and microbes scavenge the
nutrients almost as soon as they become available,
thereby immobilizing them. Such immobilization
can be disrupted by disturbances, as discussed below,

Fig. 12.7.The percentage of original nitrogen remaining in


decomposing wood in lodgepole pine forests increases before
it begins to decline; the wood slowly releases this limiting
element for the growth of other plants and microorganisms.
Adapted from Fahey (1983).

% ORIGINAL REMAINING

forms, namely, amino acids.21 In one form or another,

180
140

Nitrogen

100

Dry mass

60
20
0

12

24

36

48

60

TIME (months)

72

84

96

208 Foothills and Mountains

frees the nutrient for use by trees, plants, and microbes.

canopies, with important consequences for the abun-

The microbes in decaying wood also immobilize phos-

dance and diversity of the organisms making up the

phorus and calcium. Considering that annual nitrogen

community. Disturbances also alter nutrient cycling

inputs from fixation and precipitation are low in Wyo-

and energy and water flowseffects that are both

ming coniferous forests, as are decomposition rates, the

immediate and long term. The following sections first

nutrient-enriched decaying logs appear to be an impor-

focus on the effects of fire and insects, disturbances

tant source of nitrogen for sustained site productivity.

24

that have affected Wyomings forest ecosystems for

Silvicultural practices that remove excessive amounts

millennia, and then compare such effects to those of

of wood may be detrimental to long-term site produc-

timber harvestinga more recent and different kind

tivity, as discussed later in this chapter. Notably, only

of disturbance.

about 4 percent of a forest ecosystems nitrogen is in


living biomass. Most of it, about 90 percent, is in soil
organic matter (see fig. 12.2), with about 6 percent in

Fire and Insects

aboveground litter (including decomposing wood). Soil

The effect of fire on ecosystem processes depends on

organic matter (including associated bacteria and fungi)

the amount of heat released and the amount of vegeta-

is mineralized very slowly; therefore, only a small pro-

tion burned, characteristics that are highly variable.

portion of the nitrogen in that compartment is made

At one end of the spectrum are low-intensity surface

available each year.

fires, which historically were most common in foothill

Limiting nutrients like nitrogen are retained tena-

woodlands dominated by ponderosa pine and Douglas-

ciously once they are in the coniferous forest biomass,

fir. Such fires burn primarily litter, small trees, and

both above- and belowground. 25 One of the most

understory plants. Most canopy trees survive the fire,

remarkable adaptations of many plants is the ability to

and most forbs, grasses, and shrubs quickly regrow by

withdraw some nutrients from senescing leaves back

sprouting, depending on rooting depth and fire inten-

into stems and twigs prior to leaf fall. This process,

sity. The leaf area of the forest as a whole remains about

known as nutrient resorption, conserves as much as

the same as before the fire, and only a portion of the

half the nitrogen that might have otherwise been lost.

litter is consumed. Consequently, low-intensity surface

In this way, the plants avoid the inevitable competition

fires bring about only small changes in plant growth,

with other organisms that would occur if the nutrients

streamflow, and nutrient cycling, though such fires are

were returned to the soil. More nutrient resorption was

important for maintaining an open stand structure in

found to occur in lodgepole pine than in Engelmann

ponderosa pine and Douglas-fir forests (see chapter 11).

spruce and subalpine fir, which surely contributes to the

At the other end of the spectrum are high-intensity

ability of the pine to tolerate less-fertile soils. In gen-

crown fires, which are typical in lodgepole pine and

eral, the nitrogen concentration of forest floor litter in

spruce-fir forests. These fires kill almost all above

coniferous forests is low (less than 1 percent), a factor

ground plant parts and greatly reduce the amount of

that limits the metabolism of bacteria and fungi and

leaf area (fig. 12.8). The immediate effects are dramatic.

thereby slows decomposition. Depending on site condi-

Plant growth stops until understory plants produce

tions, complete mineralization of the leaves of lodge-

new sprouts, and evapotranspiration is greatly reduced,

pole pine is a slow process, requiring 1222 years. Boles,

because most of the leaf area has been burned. Conse-

branches, and woody roots decompose slowly as well,

quently, a much larger portion of the annual precipita-

with a 12-inch diameter tree bole requiring about 100

tion leaves the watershed as streamflow. 27 Commonly,

years for complete mineralization.

flood peaks are higher than before the fire, depending

26

on such factors as weather conditions during the snow-

Effects of Disturbances on Ecosystem Processes

melt period, as discussed previously. Also, mineral soil is


exposed when the forest floor is burned, thereby increas-

As discussed in chapter 11, both natural and human-

ing the potential for erosion.28 The actual amount of

caused disturbances thin or temporarily remove forest

erosion depends on several factors, including slope, fire

The Forest Ecosystem 209

Fig. 12.8. Grasses, sedges, and forbs sprouting from unburned roots and rhizomes 3 years after a high-severity
crown fire in a lodgepole pine forest on the northeast side
of Yellowstone Lake (the East Fire of 2003). Such plants
sequester limiting nutrients as they become available after
the fire, preventing their loss from the ecosystem. Small tree
seedlings were becoming established at the time this photo
was taken but are too small to be seen. One effect of such

fires is to mineralize the forest floor and most of the leaves,


twigs, and small branches. However, most roots, tree boles,
and large branches are not burned; and most of the organic
matter and nutrients in this unburned biomass is eventually
incorporated into the soil. In burned lodgepole forests such
as this one, about 80 percent of the large-diameter wood is
not burned, about 10 percent is converted to persistent charcoal, and the remainder is consumed.

intensity, amount of forest floor burned, water infiltra-

and total plant growth rates close to values measured

tion rate, the rate at which understory plant cover devel-

in comparable mature forests. 29 Most stands required a

ops following the fire, and whether the burn was patchy

longer time.

or uniform. The biggest erosion events usually follow

Using data from young lodgepole pine forests devel-

heavy summer rainstorms within the first year or two

oping after the 1988 fires, along with measurements

after the fire. Snowmelt is less likely to produce serious

from older forests that grew after earlier fires in Yellow

erosion than intense summer rain, even on a severely

stone National Park, Dan Kashian and colleagues pro-

burned site, because snow melts slowly and much of the

duced a model portraying the expected changes in

water percolates through the soil rather than flowing

total plant growth over a 250-year period following

over the surface.

fire or other intense disturbances. 30 Though initial

In burned lodgepole pine forests, leaf area and total

stand density makes a big difference in total productiv-

plant growth return to pre-fire levels within a few

ity, both dense and sparse stands converge on a simi-

decades. The rate of recovery depends mainly on the

lar low level of net productivity near zero by age 250

abundance of new lodgepole pine saplings, which can

years (fig.12.10). This level of net productivity does not

vary immensely, as discussed in chapter 11. In gen-

mean the trees are dead. Photosynthesis continues, but

eral, total plant growth in young stands increases with

accumulation of new plant tissues declines, for reasons

increasing tree density, although this productivity may

not yet understood.31 Lodgepole pine forests typically

be lower in stands having extreme densities (fig. 12.9).

recover more rapidly after an intense crown fire than do

Only 10 years after the 1988 Yellowstone fires, a few

any other forest types in the Rocky Mountains, except

stands of lodgepole pine already supported leaf areas

aspen forestswhich can be restored by sprouting from

210 Foothills and Mountains

7
6

12

5
4

3
2

1
0

10

10

10

10

10

Pine saplings/hectare

in the Rocky Mountains have a variety of impressive

Total LAI m2/m2

Total ANPP t/ha/yr

16

10

Fig. 12.9.Total aboveground plant growth rate, also known as


aboveground net primary productivity (ANPP), and leaf area
index (LAI), defined as square meters of leaf area per square
meter of soil surface, in relation to tree density (number of
saplings per hectare) in 10-year-old lodgepole pine forests
developing after the 1988 fires in Yellowstone National Park.
Both ANPP and LAI tend to increase with increasing sapling
density until extremely high densities of 40,000 saplings per
acre are reached. LAI in such places is almost as high as in
some mature forests. ANPP and LAI appear to level off or even
decrease in stands of even greater density (not shown here),
but the data do not permit this prediction. See fig. 12.1 for
unit conversion. Adapted from Turner et al. (2004).

mechanisms for conserving nutrients. As noted, microbial organisms, notably bacteria and fungi, sequester
them in their biomass. Also, some of the first plants to
grow back after fire absorb nitrogen and other nutrients
in excess of what they actually need for growtha phenomenon referred to as luxury consumption.35 As a result
of microbial immobilization and uptake by recovering
vascular plants, the amount of nutrients lost is less than
might be expected.36
Over a longer time frame, the nitrogen lost to a fire
in a lodgepole pine forest is typically replenished within
4070 years through local nitrogen fixation and inputs
from rain, snow, and dust.37 As the forest regrows, nitrogen and other nutrients are ever more tightly bound in
the living and dead biomass, both above- and belowground. The forest floor, including downed wood,
immobilizes the largest amounts of nitrogen, phosphorus, calcium, and magnesiumat least during the first
decade of post-fire stand development. Living biomass
appears to be the second most important factor in nutrient immobilization, especially after 6080 years. Rates
of nutrient sequestration remain positive even in very
old stands.38

the surviving aspen root system. Rocky Mountain coni-

With regard to insects, the extensive bark beetle outbreaks now sweeping across the West have important

fers lack that ability.


All kinds of fire affect nutrient cycling, with more

effects on ecosystem processes (see chapter 11 for bark

severe burns causing greater effects.32 Some nitrogen is

beetle natural history and effects on forest structure).

volatilized by the heat, but most persists in the large

By selectively killing the larger trees in a stand, beetles

amount of unburned soil organic matter. A portion of

reduce total tree growth for several years after the peak

other nutrients (like phosphorus) remains in the ash,

of an outbreak. However, the smaller trees in the stand

which becomes incorporated into the soil within the

often are not killed, and, with fewer large trees compet-

first year after the fire if not carried away by wind or

ing for water, nutrients, and light, they grow faster than

water. With more water moving through the soil profile

before the outbreak. Understory shrubs and herbs also

as a result of reduced leaf area and transpiration, the

grow faster, resulting in recovery of total stand pro-

probability of nutrient losses from the soil is enhanced

ductivity and carbon uptake and storage within a few

with subsequent nutrient gains to streams.33 After the

decades or even a few years.39 Looking at the effects of

1988 Yellowstone fires, streams flowing from several

bark beetles and other native insects on ecosystem pro-

severely burned watersheds contained elevated levels of

cesses over long time frames, some ecologists have sug-

nitrogen and phosphorus for at least 5 years. Similarly,

gested that plant-feeding insects actually help maintain

nitrate concentrations were elevated for at least 5 years

high levels of total plant growth in forests by killing

in streams draining heavily burned watersheds after the

older, slow-growing trees and freeing resources for young

2002 Hayman fire in Colorado.

individuals. This intriguing idea requires testing, along

34

Although it would seem that nutrient losses asso-

with other possible effects of insects on ecosystems.

ciated with fire, especially losses of nitrogen, might

The death of dominant canopy trees during a beetle

impair post-fire recovery, conifer forest ecosystems

outbreak can result in a substantial reduction in total

The Forest Ecosystem 211

Total NEP
g C per m2 per yr

100

High density

50
0
-50

Low density

-100
0

50

100

150

Years since fire

200

250

Fig. 12.10. Long-term trends in plant growth and carbon


dynamics in lodgepole pine forests developing after highseverity fire in Yellowstone National Park. Nearly all growth is
accounted for by trees, but the trends are based on data for all
plants in forests after increasing time since the last fire. Net
ecosystem production (NEP) is the difference between carbon
dioxide removed from the atmosphere by photosynthesis
and that returned to the atmosphere by the respiration of all
living organisms, including plants and decomposers. NEP
is initially negativethat is, respiration exceeds photosynthesisbecause of high rates of respiration by microbes and

other organisms at a time when, after the fire, water and nutrients are more readily available for decomposition and the
biomass of photosynthesizing plants is still low. Initial postfire tree density has a long-lasting influence on NEP, because
denser young forests have a higher rate of plant growth (see
fig. 12.9). Sparse young forests eventually become more dense,
reaching their maximum rates of NEP several decades later
than the forests that initially have a higher density. Both forest types eventually converge on similar low rates of NEP. NEP
units are grams of carbon per square meter per year. Adapted
from Kashian et al. (2006).

leaf area, which could, potentially, lead to changes in

the ecosystem is a short-term increase in useable nitro-

watershed hydrology.40 For example, following the

gen in the soil and increased concentrations of nitrogen

death of more than half the trees in a spruce-fir forest

in the leaves of surviving plants.44

in the Medicine Bow Mountains of southeastern Wyo-

Overall, bark beetle outbreaks lead to some inter-

ming, evapotranspiration declined and soil moisture

esting changes in hydrology and nutrient cycling, but

increased.41 That was predictable. However, a study by

changes in the quantity and quality of streamwater

Joel Biederman and his associates, working in southern

are still not well known.45 Additional research is also

Wyoming and northern Colorado, found no increase

needed to determine whether the overall effect of the

in outflow after beetle-caused mortality. The explana-

outbreaks is to cause forests to become a source or sink

tion appears to be that, although evapotranspiration

of carbonan important consideration in modern for-

was reduced by the beetle attack in the pine forest, the

est management.46 Beetle outbreaks may have cascad-

amount of sublimation from the snow increased under

ing effects on forest ecosystems,47 just as wolves appear

the more open forest canopy. Thus, no additional water

to have had in the Greater Yellowstone Ecosystem (see

was available for streamflow.

chapter 15).

42

Bark beetle outbreaks also affect nutrient cycling.


The magnitude of this effect is small compared with
the impacts of a high-intensity fire, but dead needles

Timber Harvesting Compared with Fires

falling from the canopy bring additional nutrients to

Changes in plant growth rates, hydrology, nutrient cy

the soil surface, and dead trees no longer take up nutri-

cling, and the abundance and distribution of different

ents from the soil, at least during the first few years after

plants and animals can result after timber harvesting,

an outbreak. Not surprisingly for an ecosystem where

just as they do after other disturbances. The biggest

nitrogen is a limiting factor, the surviving trees and

distinction is whether the forest was clearcut or selec-

understory plants sequester many of the nutrients that

tively harvested. Both clearcutting and crown fires kill

become available by decomposition. The net effect on

all or most trees, and it is tempting to conclude that

43

212 Foothills and Mountains

Fig. 12.12. After several generations of trees have grown


and died in a forest, every square foot of soil has had a tree
growing on it or a decomposing log laying over it. Thus, even
though evidence of decomposing logs may not be apparent on the surface (see fig. 11.17), the surface soil contains
many wood particles that contribute to soil development.
Decomposing woody roots contribute to soil development as
well. Reducing the amount of wood added to the soil by too
frequent timber harvesting could cause long-term changes in
soil structure and chemistry. Fires do not normally burn all
the wood (see fig. 12.8).

Fig. 12.11.In contrast to fire, timber harvesting removes much


of the larger wood. Leaves, twigs, and small branches are sometimes left behind as slash, but they decompose more rapidly
than the stem wood. In some places the slash is piled and
burned, reducing the amount of organic matter left on the site.

Clearcutting leaves most slash on the site, that is,


the branches, twigs, and leaves of the harvested trees
along with the bole wood of small trees that are not
removed; fire consumes most of this small-diameter
material.
An intense fire consumes much of the forest floor,

their effects are the same. Indeed, some ecological im

whereas the floor remains intact after clearcutting

pacts are similar. Both dramatically reduce leaf area,

except where the litter is disturbed by roads, various

resulting in reduced transpiration, more water moving

slash treatments, or large equipment.

through the soil, and increased streamflow. Leaching of

The amount of shade cast by dead standing trees after

soil nutrients potentially can increase with both distur-

a fire may be considerable, which can be beneficial

bances.48 Increased light and reduced competition allow

for the establishment of plant seedlings; a clearcut

shrubs and herbs to grow more prolifically after both

provides little shade if most bole wood is removed.

clearcutting and fire, which can be beneficial for herbivores, including elk and deer.
Still, there are important differences that can be enumerated as follows (figs. 12.11 and 12.12):

The differences in ecosystem structure that result


from fire and timber harvesting also lead to differences
in ecosystem processes. The intact forest floor that
remains after clearcutting continues to protect the soil

During clearcutting, most large-bole wood is re

from erosion; in fact, except on steep slopes, most ero-

moved, because wood is the desired commodity.

sion from clearcuts usually comes from the roads built

In contrast, following a crown fire, most dead boles

to access the timber, not from the harvested forest. The

remain. Even an intense fire consumes only a small

bare soil created by an intense fire is more vulnerable to

proportion of the bole wood of living trees.

erosion until plant cover is re-established by sprouting

The Forest Ecosystem 213

Fire
Coarse wood created
by disturbance
Pre-disturbance
Coarse wood

Coarse wood added


by developing stand

Fig. 12.13. Predicted long-term trends


in coarse wood after natural fire (top)
or clearcutting (bottom) in lodgepole
pine forests. Time progresses from left
to right in these diagrams. Natural
fires have occurred at centuries-long
intervals in these forests. Thus, only a
single disturbance is shown in the top
panel. The bottom panel depicts a 100year clearcutting rotation, illustrating
how repeated harvest leads to a gradual
depletion of large pieces of wood (more
than 3 inches in diameter) and possibly
to changes in the amount and characteristics of soil organic matter. Adapted
from Tinker and Knight (2004).

Initial Clearcut

Pre-disturbance
Coarse wood

Coarse wood created


by disturbance

Subsequent clearcuts

shrubs and forbs, which usually takes 14 years.49 After

the slash are released slowly during decomposition,

a fire, many nutrients formerly tied up in the living and

reducing the likelihood of nutrient losses to streamflow.

dead biomass exist in the form of ash, which can be car-

Some of the most important differences between

ried away by water and wind; some are lost in smoke as

clearcutting and fire have to do with the dead trees left

well. Warmer daytime soil temperatures resulting from

by firethe same trees that would have been harvested.

removal by fire of the cover of insulating litter also

Dead standing trees provide habitat for many organisms

stimulates soil microbial activity, including the bacteria

long after the fire.50 Also, over the long term, the addi-

and fungi that decompose dead roots and other organic

tion of wood fragments to the soil enhances nutrient

matter, thereby accelerating nutrient release. Thus, the

availability and water-holding capacity and creates an

potential exists for considerable nutrient loss after a fire.

important substrate for microbial organisms. As com-

In contrast, aside from nutrients lost by erosion along

monly noted, the character of forest soils today is par-

roads, the nutrient loss from harvesting is more likely

tially the result of organic matter additions from wood

to be less than that following an intense fire, simply

and leaves over the past 10,000 years or more (fig. 12.13).

because the harvested wood has surprisingly little nutri-

Some wood is lost to periodic forest fires, but between

ent content compared to the slash that is left behind.

fires wood-derived organic matter gradually accumu-

Unless the slash is piled and burned, the nutrients in

lates. Clearcutting interrupts these inputs, especially if

214 Foothills and Mountains

stands are repeatedly harvested at short intervals and if

vides perches for some birds, such as raptors, as well as

the slash is burned.51

a continuous supply of dead standing trees and downed

With increasing recognition of the ecological

wood in the future as these trees die and fall. The dead

importance of large wood on the forest floor and in the

trees, known as snags, also provide important habitat

soil, researchers and managers are exploring ways to

for cavity-nesting birds, such as woodpeckers, small

modify clearcutting operations to increase the amount

owls, and nuthatches.55

of wood that remains after harvest. For example, mod-

Selective cutting leaves many trees in place (fig.

eling studies suggest that simply doubling the usual

12.14) and can mimic the spatial patterns and variable

amount of slash and large wood left on a site after

intensities of insect outbreaks and low-severity fires.56

every harvest could bring long-term wood dynam-

Moderate levels of evapotranspiration are still possible,

ics much closer to the conditions associated with the

and the root systems of the remaining trees continue

natural history of these forests.52 Piling and burning

to take up nutrients and minimize erosion. One study

this material, with the intent of reducing flammabil-

in a lodgepole pine forest estimated that removing

ity or to reduce habitat for certain insects, may not be

about half the total leaf area resulted in nearly dou-

advisable.

bling the amount of water outflow, yet the outflow of

It is disconcerting to some observers that forest

nitrogen remained low. In contrast, clearcutting led

firesa natural phenomenonmight lead to more

to a nearly threefold increase in water outflow and a

erosion and nutrient loss than clearcutting, which is

sixfold increase in total nitrogen outflow. The marked

an artificial, human-caused disturbance. Such adverse


effects must be placed in the context of current conditions. Erosion from mountain forests before road
construction and similar kinds of construction was
probably less common than it is today, a time when
all erosion is viewed as a sign of poor management.
However, prior to this modern predicament, the episodic but infrequent addition of sediments and nutrients to streams could have been beneficial, in terms
of higher stream productivity and possibly higher biological diversity. Various studies have hypothesized
beneficial effects of erosion and nutrient additions to
aquatic ecosystems after fires, even to the point of suggesting that pulses in aquatic productivity might be
tied to periodic fires that burn a substantial portion
of the watershed.53 Such benefits are doubtful today
because of the numerous sources of eroded materials
in most landscapes. Exceptions are in some large roadless areas, including those designated Wilderness by
Congress.
Thus far, clearcutting has been the focus of this discussion, because it is the most intense kind of humancaused disturbance and is most comparable with crown
fire, the most intense disturbance that some consider
natural. However, less-intense selective harvesting is
often feasible. 54 It is now common to leave numerous
trees standing in what otherwise might be considered a
clearcut. Known as green tree retention, this practice pro-

Fig. 12.14. Selective harvesting, or thinning, preserves much


of the forest structure in this ponderosa pine forest in the
Black Hills. Comparable results may not be achieved in other
kinds of forests.

The Forest Ecosystem 215

differences in nitrogen outflow between thinned

ics become more frequent because of warmer winters,

and clearcut stands suggest that the surviving trees

and will the number of costly fires increase because of

in thinned stands play an important role in nutrient

longer fire seasons? What are future forests likely to look

immobilization.57

like, and how should the forests be managed to conserve

A confounding factor when selecting a silvicultural

their biological diversity? Plausible scenarios can be pro-

system for lodgepole pine forest is whether the manage

posed, given what has been learned thus far about forest

ment goal is to reduce the amount of dwarf mistletoe

ecology, the effects of recent climate change, and the

(see chapter 11), which, though a native species, is

responses of Rocky Mountain forests to previous epi-

viewed as a problem for timber production. If that is the

sodes of climate change (see chapter 2).

goal, then managers strive to cut all trees infected with

Warming has been occurring for many years and

the mistletoe. That seems simple, but the mistletoe is

is well documented (see chapters 2 and 3). Less well

difficult to see on many trees, which invariably leads to

known is whether precipitation over the mountains

some infected trees being missed and allowing mistle-

will decline. Even if annual precipitation does not

toe to spread to pine seedlings. As noted in the previ-

change substantially, warmer temperatures will cause

ous chapter, dwarf mistletoe was probably controlled in

evapot ranspiration to increase, especially in the sum-

pre-European times by periodic fires. In some places, it

mer, leading to more frequent droughts. Also, warmer

is difficult to use timber harvesting to accomplish the

winters will cause more precipitation to fall as rain

same objectiveexcept by clearcutting. Another con-

rather than snow, and a shallow snowpack combined

cern with partial cutting of lodgepole pine, Engelmann

with warm spring temperatures will lead to earlier

spruce, and subalpine fir forests is that the residual trees

snowmelt, earlier drying of fuels, a longer fire season,

are sometimes vulnerable to windthrow.

and a longer period for trees to be water stressed. Very

Regardless of whether the forest is subjected to par-

likely, climate change will lead to more frequent insect

tial or total tree deathwhether by burning, cutting, or

outbreaks; more frequent large fires; and reduced water

insect epidemicsthe lost nutrients are a small portion

supplies for towns, industry, and agriculture. Forest

of the total nutrient capital. The leaching of nutrients is

regeneration will sometimes fail if extended droughts

elevated for perhaps 10 years, possibly more, but the dis-

become more common58 and if more frequent fires

turbances do not slow nutrient inputs. Most likely the

occur. Most forest types may shift upward in eleva-

lost nutrients are replaced in a few decades at the most.

tion, as the alpine climate becomes favorable for tree

Soil erosion is a more serious concern. Still, with proper

growthresulting in less total forest cover because

precautions it should be possible to ensure the continu-

there is less land area at higher elevations.59 Consider-

ation of soil development and retention. Though hav-

ing the current climate requirements of the prevalent

ing significant impacts at the time they occur, both

forest trees today, U.S. Forest Service scientist Gerald

crown fires and clearcuts are infrequent events. A big-

Rehfeldt and his colleagues concluded that, by the end

ger concern is the proportion of a watershed or forested

of the twenty-first century, forests are likely to cover

landscape affected during a period of time.

less area in the western United States than they do


now.60 Today, forests cover only about 15 percent of

The Future of Mountain Forests

Wyoming (see fig. 1.5)


What can be done about the prospect of reduced

Many forests in the region show the imprint of a century

water supplies? Knowing that transpiration is reduced

of intensive use, having been fragmented by roads, tim-

and water outflow enhanced in forests having less leaf

ber harvesting, homes, and other developments. Others

area, as discussed above, some planners have suggested

have been altered only slightly, such as in roadless areas.

that water supplies could be augmented by increasing

Everywhere, however, big changes loom because of cli-

the harvesting of mountain forests.61 Although possible,

mate change. Will warming affect precipitation patterns

substantial increases could be produced only from forests

in ways that reduce water availability for tree growth

that receive the largest amounts of precipitation, that is,

as well as downstream users? Will bark beetle epidem-

forests at the highest elevations, and such increases in

61

Spring-Summer
Temperature

100

59

Wildfire Frequency

216 Foothills and Mountains

57

0
1970

1975

1980

1985

1990

1995

2000

55

Fig. 12.15.Trends in wildfire frequency and average springsummer temperature (March through August) in the western
United States since 1970. Fire frequency is the number of fires
greater than 1,000 acres in size per year. Temperature and

fire frequency are correlated, and both increased beginning


around 1985. Climate models suggest these trends are likely
to continue during the twenty-first century. Adapted from
Westerling et al. (2006).

water would be seen only in years of average or above-

Have Past Fires Increased the Probability of


Beetle Outbreaks?

average precipitation, which may become infrequent.


Because the regrowth of high-elevation forests is slow
and such forests are highly valued for other reasons
wildlife, recreation, and aestheticsintensive harvest
may not be a socially acceptable strategy for increasing
water supplies. Other approaches, such as improving
water conservation measures and augmenting water
storage facilities, may be more effective.
With regard to fires, an upsurge of large fires during the past 25 years has already occurred across most
of western North America (fig. 12.15). In addition,
extensive bark beetle outbreaks are now a continental
phenomenon.62 Several factors may be responsible for
this recent upswing in forest disturbance. One is the
extensive cover of dense, mature conifer forests that
developed during the twentieth century, when fires and
other natural disturbances were relatively infrequent,
and when human-caused disturbances (such as timber
harvest) affected only a small portion of the forested
land. Dense, mature forests generally are more susceptible to fire and insect outbreaks than are younger
or more open forests. It is tempting to attribute such
problems to inadequate forest management, that is, too
much fire suppression or too little timber harvesting.
However, the increase in fire and bark beetle activity

For many years ecologists thought that fire-injured trees


provided loci for subsequent mountain pine beetle outbreaks into surrounding unburned forests. The results
of one study in the Greater Yellowstone Ecosystem re
vealed that the number of new adult beetles in the next
generation was greater in trees that sustained moderate
fire injury than in severely injured or uninjured trees.64
However, moderately injured trees are usually scarce
after fire in lodgepole pine forests, as most trees touched
by fire are severely injured and die. Few new adult
beetles emerge from the severely injured trees because
of various factors, including damage to the inner bark
tissues on which the beetles depend.
Drought or other nonfire stresses are more likely
initiators of mountain pine beetle outbreaks, simply
because they cause healthy trees to be more susceptible
to beetle attack without damaging the food resource
used by the beetle larvae. Also, in contrast to the mountain pine beetle, spruce beetle outbreaks are known
to begin in windthrown timber or large-diameter logging slashthough drought or other conditions that
stress living trees are necessary to sustain a widespread
outbreak.65

is almost ubiquitous, occurring in forests with a great


variety of management histories throughout most of
the western and northwestern states and up into Alberta
and British Columbia. Climate change is the most likely
explanation;63 current climate models suggest that conditions favorable for beetles and fire will become more
frequent in the future.

Do Beetle Outbreaks Increase the Probability of Fire?


After a beetle epidemic, forests are mostly reddishbrown or gray and they appear alarmingly flammable.
Research suggests that this association is not as simple
as it seems. Jeffrey Hicke and his colleagues from the
University of Idaho developed a model that character-

The Forest Ecosystem 217

izes general fuel conditions and potential fire behavior

extremely dry, everything will.68 Although beetle-

in conifer forests during four stages of an outbreakthe

affected forests appear highly flammable for several

green stage early in the attack; the red stage, when the

years after infestation, green forests can burn just as

trees are covered by dead needles; the gray stage, when

intensely under hot, dry weather conditions.

the dead needles have fallen but most of the dead tree

Most forest fires occur during relatively extreme

boles are still standing; and the old stage, when the dead

weather conditions, especially in lodgepole pine and

trees have fallen and a new overstory is developing.66

spruce-fir forests. Consequently, it is not surprising that

Their model incorporates data on surface fuels, which

the influence of beetles on flammability has been small

support a surface fire of varying intensity; crown bulk

or undetectable. One study of the 1988 Yellowstone fires

density or total mass of needles and small twigs on the

found that forests were more likely to have burned in

trees, which can support the burning of tree crowns;

1988 where high beetle-caused tree mortality occurred

and ladder fuels, which are small living trees that can

15 years earlierthough only about 10 percent more

carry fire from the surface fuels into the canopy. Two

likely.69 An examination of all fires in northwestern

types of crown fire are possible: torching (the burning

Colorado and southern Wyoming during the past cen-

of some individual tree crowns) and active crown fire

tury found that forests where bark beetle outbreaks had

(in which the flames spread from tree crown to tree

occurred previously were no more likely to have burned

crown and through most of the forest canopy, killing

with high severity than forests with no previous his-

the majority of trees).

tory of beetle activity.70 In contrast, years with large

Hicke and colleagues found no substantial change

high-severity fires were all drier than years without

in surface fuels or canopy bulk density during the red

fires. Overall, drought and windnot beetlesare most

stage, but the dry, dead needles at this stage increase the

highly correlated with high-severity fires.

probability of torching and active crown fire. Crown

The implication of these recent studies is that large

bulk density is greatly reduced during the gray stage,

fires and extensive insect outbreaks tend to occur

which reduces the likelihood of active crown fires. How-

together, not because one causes the other but because

ever, surface fuels increase during the gray stage, lead-

both respond to similar climatic conditionswarmer

ing to a potentially more intense surface fire, which in

winter and summer temperatures and earlier melting

turn may produce torching of individual tree crowns.

snowpack. These conditions create a longer fire sea-

In the old stage, the fallen boles of the beetle-killed trees

son and greater drought stress for trees in late summer.

contribute to a potentially intense surface fire. More-

Stressed trees are less able to defend themselves against

over, the ladder fuels that have developed at this stage

bark beetles, and the beetles themselves are better able

increase the potential for crown fires.

to survive in the winter, because long cold periods are

Two primary caveats are necessary when consider-

less frequenta convergence of conditions that favor

ing this conceptual model. First, there is disagreement

both fire and beetle disturbance and that is likely to

about many of the models interpretations, partly

become more frequent in the future.71

because bark beetle effects are variable and therefore


workers in different settings will reach somewhat different conclusions.67 Second, although the conceptual

Some Management Options

model focuses on the effects of changes in fuel condi-

In large roadless areas, such as the backcountry of

tions, weather conditions at the time of a fire are as

Yellowstone National Park, managers often allow fires

importantor more importantthan subtle differ-

and insect outbreaks to spread without interference.

ences in fuels. When fuels are dry, temperatures are

Few resources are threatened and such disturbances

high, and winds are strong, any of the four stages

bring about ecological benefits. Native plants and ani-

are easily burned. Indeed, the patterns described in

mals are adapted for such events; the ecosystems are

the model may be relevant only during intermedi-

resilient.72 This is true even if there are more fires in

ate weather conditions, that is, when not too wet and

coming decades. Studies of charcoal and pollen pre-

not too dry. When wet, nothing will burn and when

served in lake sediments reveal that the Yellowstone

218 Foothills and Mountains

country burned more frequently at times in the past,

wind or oriented parallel to the direction of expected

notably about 10,000 years ago, when summers were

fire spread. Significantly, the protection afforded by fuel

warmer than today, yet lodgepole pine apparently was

breaks of this kind are often compromised by glowing

as abundant at that time as it is today.73 The results of

embers, which can be transported by the wind for sev-

this research suggest that forests of lodgepole probably

eral hundred yards and often ignite new fires. This is a

will persist in spite of increased burning and insect

common cause of fire spread on windy days.

activity, though a recent study suggested that the future

Another landscape-level approach is to prohibit

climate in some areas may be unsuitable for this pine,

or discourage home construction in areas of greatest

irrespective of such disturbances.74 Generalizations are

wildfire risk. This might be best accomplished through

difficult; most likely, the dominant factors bringing

disincentives, such as higher property tax rates and

about change will vary from place to place.

insurance premiums, or by county restrictions that

But not all Rocky Mountain forests are in roadless

reflect the relative fire risk at a given location and the

areas. The number of people building homes in moun-

costs to the community as a whole. Access for fire sup-

tain forests has increased greatly over the past two

pression is often presented as a justification for building

decades because of the appeal of living in such an envi-

more roads, but most areas in the foothills and adja-

ronment. Timber harvesting also is an important activ-

cent mountains already have extensive road networks.79

ity in much of the West. Large forest fires are of great

The negative effects of additional roads are discussed in

concern in these settings. Two general approaches can

chapter 18.

be used to reduce the threat of fire. The first is local, that

How should managers deal with bark beetle out-

is, reduce the density of trees, shrubs, and other flam-

breaks? If an outbreak is not already in progress, man

mable vegetation in the immediate vicinity of a house or

agers can reduce the likelihood of one occurring locally

other structure and build the structure itself with non-

by using standard techniques of good forestry, that is,

flammable materials.

Many homes that implemented

by thinning and spacing trees for optimal growth and

the Firewise recommendations of the U.S. Forest Ser-

the maintenance of healthy trees that are capable of de

vice have escaped serious damage in recent fires.76 Fuel

fending themselves against bark beetles. 80 Making forest

reduction also can be accomplished in forests further

stands more resistant to bark beetles in this way is prob-

removed from vulnerable structures. The 2000 National

ably most appropriate on public and private lands where

Fire Plan directs federal agencies to conduct thinning

timber production is an important management goal

and prescribed fire treatments in this wildland-urban

or in campgrounds and communities where the death

interface; however, this mandate has proved challeng-

of many large trees would impair local aesthetics and

ing to accomplish, in part because most of the interface

create the hazard of falling trees. Such intensive, active

is on private, not public, land.77

management would be highly controversial in roadless

75

The second approach for reducing fire risks is applied

areas and in the backcountry portions of national parks.

on a broader scale. The objective is to break up continu-

Once a beetle outbreak is well under way, stopping

ous fuels and create places where fire spread and inten-

it does not appear to be an option at the present time.

sity can be reduced and where suppression crews can

Regardless of management, the outbreak will continue

attack fires effectively when that seems necessary. Forest

until bitter cold temperatures at the right time kill

fire scientists have identified optimal sizes and position-

the beetle larvae wintering under the tree bark, or the

ing of fuel reduction sites to achieve the greatest pro-

beetles simply run out of suitable host trees.81 Individual

tection of vulnerable resources for a given investment

high-value treessuch as those in campgrounds, town

in fuel reduction.78 For example, large fires are often

parks, or yardscan be protected from beetles even

driven by high winds coming from more-or-less pre-

during an outbreak by repeatedly applying appropriate

dictable directions. Therefore, rectangular fuel breaks

insecticides, but, over large areas, that solution has been

of sufficient size placed upwind of a community and

prohibitively expensive and problematic to date.

oriented perpendicular to the prevailing winds could be

Alternatives are available for handling the trees killed

more effective than similar fuel breaks located down-

by the beetles, such as removing them from around

The Forest Ecosystem 219

homes, campgrounds, and along roads and trails to

growing; other clones appear to be dead or have very

prevent damage or injury when the dead trees fall. Un

few sprouts.84

fortunately, cutting and burning beetle-killed trees after

At the same time that aspen appeared to be declin-

the needles have turned red will do nothing to slow a

ing at low elevations, new aspen seedlings that became

mountain pine beetle outbreak, because the new cohort

established in burned forests after the 1988 Yellow-

of beetles has already emerged from the tree by the time

stone fires were growing best at higher elevationsin

the needles dry and begin to fall. Removing beetle-killed

some places higher than the pre-1988 range of aspen

trees is also questionable as a fire mitigation method,

in Yellowstone.85 The same pattern has been seen after

because, as discussed, beetle-caused mortality does not

recent fires in the Canadian Rockies. In Colorados San

appear to increase fire hazard.82 After 2050 years, when

Juan Mountains, a comparison of historic photos from

the dead trees have fallen and the surviving small trees

the early 1900s with recent photos of the same land-

are again forming a forest canopy, the heavy fuel load

scape revealed new aspen clones at the upper treeline,

will have the potential to support an intense fire. How-

apparently the result of seedling establishment even

ever, it is not known whether future fires in forests with

without any major disturbance by fire.86

many beetle-killed trees will actually be more intense

The pattern of aspen mortality at lower elevations

than fires in a green forest where that has not happened.

and new recruitment at higher ones suggests that aspen

Hence, at this time no clear guidelines exist for whether

is already responding to the climate changes of the past

trees killed by mountain pine beetle should be removed

quarter-century. Such trends are likely to continue,

from locations that are distant from homes, roadsides,

with some low-elevation clones dying as climate con-

and other obviously vulnerable resources. Doing so may

ditions become less favorable there. Forest ecologists

be more expensive than can be justified, considering that

and climate modelers have worked together to project

the next fire or beetle outbreak will probably be influ-

the future distribution of suitable habitat for aspen and

enced more by weather patterns than by fuel conditions.

other plants in western North America. Their approach

Again, however, the spruce beetle presents a different

is to map the current range of aspen and to then

problem, and the removal of dead trees may be advisable

describe the extent of variation in climatic conditions

when they are a threat.

that exist throughout the current range, thereby creating a bioclimatic envelope. They then use climate models

Aspen Forests: Bellwether of Future Changes?

to identify specific locations where suitable conditions


for aspen will be located in the future, assuming vari-

Quaking aspen is one of the most highly prized trees in

ous scenarios for climate change. For aspen, two major

western North America, primarily for the biological and

expectations emerge from these initial efforts: (1) the

aesthetic diversity it adds to the landscape. Thus, it was

total extent of suitable habitat will likely decrease dur-

especially alarming when large aspen trees began dying

ing the twenty-first century, although aspen will persist

in 2003 and continued to do so through 2007, mostly

in small pockets throughout most of its current range;

in Colorado and Utah but also in Wyoming and other

and (2) new groves will most likely become established

western states (fig. 12.16).83 This widespread phenom-

at higher elevations (up to 1,000 feet higher than where

enon has been called sudden aspen decline (or SAD).

aspen grows now).

For a time the cause of decline was unknown, but now

Similar studies have come to about the same conclu-

it appears that the severe drought from 2000 to 2002

sion for most other Rocky Mountain trees.87 Notably,

coupled with unusually high temperatures was the trig-

because there is less land area at higher elevations, and

gering factor. The greatest aspen decline has taken place

much of it lacks soil (such as on cliffs and talus slopes),

on south-facing slopes at lower elevations and on soil

the area occupied by each species will be reduced. In

types with low water-holding capacity, two observa-

Wyoming, the amount of forested land could decline

tions that are consistent with the drought hypothesis.

from about 15 percent at the present time to 78 per-

The root systems and some small trees of many clones

cent. There could be exceptions to this trend for spe-

survived and are now producing new sprouts that are

cific species. For example, Douglas-fir may be able to

220 Foothills and Mountains


Fig. 12.16. Sudden aspen decline affected aspen
stands in various places throughout the Rocky
Mountain region during 20012005. In some
stands, most large stems diedthough the root
systems did not. New sprouts are expected to
produce a new forest if they are not subjected
to another severe drought or to chronic excess
browsing by ungulates (see chapter 15). This
photo was taken on the Uncompahgre Plateau in
western Colorado. Elevation 9,500 feet. Photo by
Dan Binkley.

move up and across the expansive Yellowstone Plateau

mixes of species. Forests are likely to be more frag-

from the surrounding foothills where it now grows, as

mented in the future (a topic that is considered further

the cool climate on the Plateau becomes warmer. That

in chapter 18), and extensive tracts of interior forest

would happen, however, only if Douglas-fir can tolerate

may be less common than they are now. Meadows and

the relatively infertile rhyolitic soils found there.

parklands may become more common, just as they did


during the warm and dry intervals of the past 10,000

In sum, a cadre of scientists and managers is develop-

years. It seems prudent for managers to conserve native

ing procedures for predicting, detecting, mitigating,

species where possible, take actions to minimize the

and adapting to the ecological changes that can be

number of invasive species, and consider how the eco-

expected.88 The forests that future generations enjoy

system services provided by forests in the past can be

may cover less total land area and be composed of novel

extended into the future.

Mountain Meadows
and Snowglades

Chapter 13

Sometimes referred to as parks, meadows add diversity to

Higher on the mountain, Idaho fescue and tufted hair-

the mountain landscape (fig. 13.1). The plants are highly

grass are more common (see chapter 10). Wet meadows

variable due to differences in elevation, wind exposure,

are found at all elevations along streams and in depres-

water availability, and soil characteristics, but often they

sions, where melting snow provides a source of water

include such colorful species as arrowleaf balsamroot,

sufficient to support shrubby cinquefoil and mountain

Indian paintbrush, and lupine (table 13.1). Such plants

silver sagebrush in addition to various sedges and grasses

attract numerous butterflies, along with other animals

(see chapter 5).2 Meadows with intermediate moisture

that benefit from herbaceous plants. At lower elevations,

availability support the greatest diversity of plants.

the plants are similar to those in dry foothill grasslands,

A frequently asked question about mountain mead-

with bluebunch wheatgrass a characteristic species.

ows is why they have no trees. The answer depends on


Fig. 13.1. Mountain meadow
in the Wyoming Range west
of Big Piney. Mountain big
sagebrush and Idaho fescue
are common, with willowdominated shrublands in
the riparian zone. Arrowleaf
balsamroot and lupine are
abundant in the foreground.
The forests in this area are
dominated by Engelmann
spruce, lodgepole pine, and
subalpine fir, with whitebark
pine near the alpine treeline.
Elevation 8,600 feet.

221

Table 13.1. Some characteristic plants of subalpine meadows in Wyominga

Common name

Latin name

Sagebrush
meadows

Dry subalpine
meadows

Mesic subalpine
meadows

GRASSES
Alpine bluegrass

Poa alpina

Cusicks bluegrass

Poa cusickii

Muttongrass

Poa fendleriana

Nodding bluegrass

Poa reflexa

Pattersons bluegrass

Poa abbreviata ssp. pattersonii

Sandberg bluegrass

Poa secunda

Timberline bluegrass

Poa glauca ssp. rupicola

Alpine timothy

Phleum alpinum

Idaho fescue

Festuca idahoensis

Rocky Mountain fescue

Festuca saximontana

Oatgrass

Danthonia spp.

Prairie junegrass

Koeleria macrantha

Slender wheatgrass

Elymus trachycaulus

Thickspike wheatgrass

Elymus lanceolatus

Spike trisetum

Trisetum spicatum

Tufted hairgrass

Deschampsia caespitosa

SEDGES
Curly (rock) sedge

Carex rupestris

Dunhead sedge

Carex phaeocephala

Ebony sedge

Carex ebenea

Hoods sedge

Carex hoodii

Needleleaf sedge

Carex duriuscula

Northwest Territory (beaked) sedge

Carex utriculata

Obtuse sedge

Carex obtusata

Sheep sedge

Carex illota

Polygonum bistortoides

Alpine sagebrush

Artemisia scopulorum

Prairie (fringed) sagewort

Artemisia frigida

Arrowleaf balsamroot

Balsamorrhiza sagittata

Ballhead sandwort

Arenaria congesta

FORBS
Alpine bistort

Creeping sibbaldia

Sibbaldia procumbens

Elephanthead lousewort

Pedicularis groenlandica

Field (mouse-ear) chickweed

Cerastium arvense

Fireweed

Chamerion angustifolium

Indian paintbrush

Castilleja spp.

Mules ears

Wyethia amplexicaulis

Pale agoseris

Agoseris glauca

Mountain Meadows and Snowglades

Sagebrush
meadows

Dry subalpine
meadows

Mesic subalpine
meadows

Common name

Latin name

Purple locoweed

Oxytropis lambertii

Ross avens

Geum rossii

Silvery lupine

Lupinus argenteus

Sulphur-flower buckwheat

Eriogonum umbellatum

Tall fringed bluebells

Mertensia ciliata

Tufted fleabane

Erigeron caespitosus

Western yarrow

Achillea millefolium var.


occidentalis

White locoweed

Oxytropis sericea

White marsh marigold

Caltha leptosepala

Artemisia tridentata ssp. vaseyana

SHRUBS
Mountain big sagebrush
Silver sagebrush (mountain)

Artemisia cana ssp. viscidula

Shrubby cinquefoil

Dasiphora fruticosa ssp. floribunda

A dash indicates that the plant is absent or uncommon.

local conditions. Some meadows occur on fine-textured

common, surface-soil texture is coarser. The presence

soils along rivers and streams, where the absence of trees

of a distinct stone line under the fine-textured surface

has been attributed to the soils remaining too wet for

soil in the meadow and its absence in the forest led

too long in the summer.3 Elsewhere, competition from

Doering and Reider to conclude that the meadow origi-

grasses and forbs may be too great. Other meadows are

nated after a severe fire or windstorm, perhaps several

found where trees are excluded because the soils are not

thousand years ago. Subsequent wind erosion removed

favorable for tree seedling establishment (fig. 13.2), or

much of the exposed fine material but left a layer of

where fires have occurred at high elevations and the re-

stones (fig.13.4). Gradually, new deposits of fine, wind-

establishment of new trees is slow. Transitions between

blown material covered the park. The dryness of the

meadows and forests in upland settings can be difficult

park probably is caused by two factors: (1) the diffi-

to explain, as soil differences often are not obvious.4

culty of water percolation through the shallow stone

Perhaps the most-studied meadow in the Rocky

line, with most water held in the shallow, fine-textured

Mountain region is Cinnabar Park in the Medicine Bow

surface soil, where it is readily evaporated; and (2) the

Mountains (fig. 13.3). Early reports suggested that this

strong winter winds that greatly reduce snow accumu-

dry meadow is slowly moving downwind, with trees

lation. Blowing ice and snow also damage the leaves

invading on the upwind side (where more snow accu-

and stems of young tree seedlings, leading to further

mulates) and older trees dying on the downwind edge.

mortality.6 Plant water stress in the meadows is sug-

However, soil scientists William Doering and Richard

gested by the obvious wilting of many plants during

Reider concluded in 1992 that the park was stable.

clear summer days.

They suggested, after detailed analysis of soil profiles,

Subalpine meadows also persist where summer frosts

that trees were absent in the meadow because of a shal-

are frequent. This seems counterintuitive, because trees

low, 6-inch accumulation of fine-textured surface soil

are found nearby. However, because of direct exposure

that provided a better environment for grasses and

of the meadows to cold skies at night, frosts at the soil

forbs. In the adjacent forest, where tree seedlings are

surface can occur on a third or more of summer nights.

223

Fig. 13.2. Large mountain meadow in the Bighorn Mountains, located where shales in the Gros Ventre Formation are
exposed. Shales lead to the development of fine-textured
soils that are not favorable for tree seedling establishment.
Common plants in the meadow include fringed sagewort,
hoary balsamroot, Idaho fescue, junegrass, prairie smoke,
silvery lupine, starry cerastium, wild geranium, and yarrow.

Aspen groves occur in the ravines. Forests here are dominated


by lodgepole pine and limber pine and are found on the
adjacent, coarse-textured soils derived from sandstones in
the Flathead Formation. Similar meadows exist where finetextured sedimentary rocks are exposed on the surface, such
as on the east slopes of the Wind River Mountains. Elevation
7,400 feet.

Fig. 13.3.Cinnabar Park, a


dry meadow in the Medicine
Bow Mountains, is located
on a high plateau. Strong
winds blow the snow off the
meadow, where it forms drifts
on the leeward side of the ribbon of lodgepole pine on the
left. Often the snow persists
until August, preventing
small trees from becoming
established. Canby bluegrass,
Idaho fescue, junegrass, sheep
fescue, and starry cerastium
are common in the dry
meadow; tufted hairgrass and
other wet meadow species
are found in the snowglade.
Elevation 9,600 feet.

Mountain Meadows and Snowglades

Fig. 13.4. Soil characteristics across Cinnabar Park. The stone


line probably formed during a period when fine soil particles
were eroded from the high plateau. New fine material gradually

accumulated to a depth of approximately 6 inches (15 cm), but


the establishment of tree seedlings in the dry meadow is still
difficult. See fig. 13.3. Adapted from Doering and Reider (1992).

Tree seedlings are easily killed by a combination of frost

topography, and the chance establishment of a small

damage and reduced rates of photosynthesis the next

tree or shrubwhich initiates snow deposition on its

day. In contrast, summer freezing is rare under the for-

leeward sideone or more bands or ribbons of trees are

est canopy, if it occurs at all. Two other factors limiting

sometimes found in the vicinity of snowglades.

tree seedling establishment could be excess competition

The ribbon forestsnowglade pattern takes at least

for tree seedlings with established herbaceous plants

two forms in the region.10 Commonly, there is just one

and the kind of vertical discontinuity in soil texture

band of trees and one band of meadow on the lee side

observed at Cinnabar Park.7

of a more extensive meadow, such as at Cinnabar Park


where the late-spring snowdrift may be 30 feet deep or

Snowglades and Ribbon Forests

more (see fig. 13.3). The initial origin of such meadows


is not easily explained. Dwight Billings, Duke Univer-

Meadows created by late-lying snow are known as snow-

sity ecologist, suggested in 1969 that forest fire near

glades. The soil is covered, cold, and wet for too long and

timberline . . . changes the snowdrift pattern enough

the growing season is usually too short for tree seed-

that trees in an unburned area to the lee of the burn

lings to become established.8 Also, deep snow that per-

are killed by late-lying snow during the summer. The

sists until midsummer favors the growth of molds that

dead trees are replaced by a wet type of snowglade

can kill buried branches and young trees, for example,

meadow.11 Support for this hypothesis comes from the

the blackfelt snowmold (Herpotrichia nigra) (fig. 13.5).

observation that huge snowdrifts form on the lee side of

In addition, pocket gophers that feed on tree roots and

some clearcuts, usually 1020 feet into the uncut forest,

bark are sometimes common under the snow.

and the effect of the drift is to reduce tree vigor. Dead

Tree seedlings, however, also have difficulty becom-

standing trees can be found even though the upwind

ing established where snow accumulation is low, in part

clearcut is less than 20 years old. Over a period of 500

because of insufficient moisture. As important may be

years or more, all remnants of the original forest could

insufficient protection from abrasion by blowing ice

disappear.

and snow. Working in the Medicine Bow Mountains,

In other situations, usually near upper treeline, sev-

Stephan Httenschwiler and William Smith found that

eral bands of forest and meadow may occur in sequence

tree seedlings were favored in places where snow depth

(fig. 13.7).12 Again, the role of wind and snow drifting

was at least 18 inches but no more than 56 inches (fig.

is obvious, because the bands are perpendicular to the

13.6).9 Due to the interaction of drifting snow, micro-

prevailing westerly winds and the deepest snow occurs

225

226 Foothills and Mountains

in the glades.13 The establishment of the first trees that


lead to the ribbon forest pattern is still a puzzle to be
solved (but see chapter 14).
A similar phenomenon has been observed where
aspen occur in doughnut-shaped or U-shaped groves
that are sometimes referred to as aspen atolls (see
fig.10.13). Commonly surrounded by sagebrush or subalpine meadows, the groves probably developed because
of the snow-fence effect of a few trees that by chance
became established. The original trees created the
only sites where new seedlings or sprouts could obtain
enough water to survive. Eventually, the clone expands
around all or most of the snowbank, such as south of
Rawlins on the north end of the Sierra Madre. Engelmann spruce and subalpine fir develop similar groves
in northwestern Wyoming, but at higher elevations.14
Trees cannot survive in the center, because the snow
persists too long, but the snowdrift provides a dependable supply of moisture for the trees on the perimeter.
More often seen are ridgetop ribbon forests. In this case,
deep snow accumulates on the lee side of ridges, maintaining a linear snowglade just above trees that benefit
from the meltwater and, consequently, are able to grow
where they might not otherwise occur.

25

The forest atoll or ribbon forestsnowglade pattern


may characterize only a small portion of the mountain
landscape, but the pattern is of considerable significance
economically, as they are a naturally developed high-

` ribbon _

20
15

W- and SW-winds

10
5

SW

NE

snow

20
15

30

20

10

Distance (feet)

none

none

none

10
none

Seedlings / ft2

Height above ground (feet)

Fig. 13.5. Blackfelt snowmolds can kill branches and small


trees that remain covered with snow late into summer, such
as in some snowglades. This photo shows an infected lodgepole pine sapling. The lower needles are compressed by the
black or brown fungal filaments (hyphae).

10

20

Distance (feet)

30

Fig. 13.6.Common pattern


in a subalpine meadow near
treeline, near 11,000 feet,
with ribbons of mature
Engelmann spruce and
subalpine fir (see fig. 13.7).
Note how the trees affect
snow drifting and how
moderate snow depths appear to favor tree seedling
establishment. The seedlings observed in this area
were 110 years old and
may not grow to maturity.
Snow accumulation depths
and spatial distribution
vary according to tree
height and the width of the
ribbon of trees. Adapted
from Httenschwiler and
Smith (1999).

Mountain Meadows and Snowglades

higher in the mountain meadows because of the variety


of plant communities that occur up and down mountain slopes.15 Essentially, there is a wave of new plant
growth that the animals can follow, from the lowlands
in the spring to the high mountain meadows later in
the summer. Because of topographic diversity and different amounts of snow accumulation, a supply of green
forage almost always exists nearby.
Livestock managers have learned that the initiation
and duration of grazing is important to consider when
developing management plans. Ranchers prefer to rest
their low-elevation rangelands in the spring, when
plants there are actively growing and are most susceptible to damage from grazing. Consequently, they want
to move livestock onto the mountain meadows as soon
as possible. The trick is to move the animals at a time
when the soils are not so wet that damage from trampling occurs. Moreover, just as in the lowlands, the
mountain plants are subject to stress if grazed excessively in the early stages of their growth.16
Today, the grazing of mountain meadows is restricted
in many areas to a shorter summer period than in the
early 1900s. Fewer animals are permitted on federal
Fig. 13.7. At high elevations with strong winter winds, such
as in this area north of Medicine Bow Peak, parallel ribbons
of trees and snowglades sometimes develop. The trees are
predominantly Engelmann spruce and subalpine fir. See fig.
10.13 for a photo of an aspen atoll, caused by the effects of
blowing snow at lower elevations. Photo by William K. Smith.

land, and often a shepherd must be employed to keep


the animalsespecially sheepfrom grazing specific
areas excessively. As a result, some ranchers now find
it uneconomical to move their livestock to the mountains for such a short time, often midsummer to early
fall. Also, the amount of land area in meadows has
declined over the years in some areas, owing to fire sup-

water-yield ecosystem, in which most of the snow accu-

pression and the subsequent expansion of trees. This

mulates where there is comparatively little use of that

trend might be especially apparent in mountain ranges

water by plants. Not surprisingly, a common practice of

that were subject to considerable burning or tree cut-

watershed hydrologists desiring to increase streamflow

ting in the late 1800s and early 1900s. The best example

is to create small openings in the forest, where snow

in Wyoming may be the Sierra Madre, where extensive

can accumulate in much the same way (see chapter 12).

timber harvesting in conjunction with copper mining


created large areas of open rangeland. Now much of this

Livestock Grazing on Mountain Meadows

area is again forested.


Overall, mountain meadows are widely appreciated

Mountain meadows have long been an important com-

for the benefits they provide in terms of livestock for-

ponent of the summer range for elk and deer, and for a

age and water yield. As important, the edge between the

century or more many ranchers have come to depend

forest and meadow provides habitat for many birds and

on such meadows for supplemental livestock forage.

mammals. Trees have invaded some meadows, but only

Using the meadows in this way during the summer pro-

rarely are forests converted to meadows after fire or tim-

vides a period of rest for rangelands in the valleys below.

ber harvesting. Shifts in the location of forest-meadow

The carrying capacity for both wildlife and livestock is

boundaries may be an indication of climate change.

227

228 Foothills and Mountains

Forest Expansion into Meadows

Response to Climate Change

Although some meadows appear to be stable features

Predicting the effects of warming (or cooling) on plant

of the landscape, others are being invaded by trees.

growth in meadows is currently hampered by un

As noted, the origin of such meadows is often difficult

certainties about how water availability will change

to explain, but the presence of young trees suggests a

under future climates. If the annual precipitation

change of one or more environmental factors over time.

increases or remains about the same, then concomitant

The factors mentioned most often are less frequent fires,

higher temperatures will create drier conditions, because

which enables trees to grow where they could not sur-

the water will evaporate more rapidlyoften before it

vive when fires were more frequent; grazing by domestic

can enhance plant growth. The annual growth of many

livestock, which can favor trees by reducing competi-

plants would be slower if summer drought stress were

tion from herbaceous plants; and climate variation.

to begin earlier. If summer precipitation increases to a

Meadows that are otherwise apparently stable could

level that more than compensates for increased evapo-

also be undergoing very slow reforestation following a

transpiration, then a warmer climate could result in a

severe disturbance that occurred long ago. An example

longer growing season, depending on whether the addi-

of slow invasion is at high elevations on parts of Libby

tional water comes as snow or summer rain. If summer

Flats in the Medicine Bow Mountains (see chapter 14).17

rainfall were to increase sufficiently, then the growing

In considering the effects of climate variation, sev-

season for many mountain plants could be lengthened,

eral scenarios are plausible. First, consider what might

and the annual production of new plant biomass would

happen if the climate were to become cooler. Evidence

increase. Alternatively, more snow may not have much

suggests that tree establishment after such disturbances

effect on plant growth, because even low snowfall years

as fire would be slowed. For example, pollen data from

most likely would produce enough meltwater to saturate

the Wind River Mountains indicate that forests gave

the soils. The additional water could augment stream-

way to meadows about 3,000 years ago during a period

flow and reservoir levels but often would not increase

of cooling.18 Nocturnal summer frosts at this time

plant growth in the meadows.

probably became frequent enough to restrict tree seed-

Of course, not all meadows and plant growth forms

ling establishment as the older conifers died, thereby

would respond in the same way. Ecologist Diane

enabling herbaceous plants to dominate instead after

Debinski and her colleagues studied dry, moist, and

fires or windstorms. The long-term persistence of pine,

wet meadows in the Yellowstone area during a transi-

spruce, and fir depends on seedling establishment, as

tion from very wet conditions (19971999) to very dry

these trees lack the ability to sprout from stumps or

ones (20002007).21 The amount of plant cover gener-

roots.

ally declined, especially in dry meadows, but changes

In contrast, if warming occurs, then summer frosts

in forb and shrub cover varied considerably. Shrub

at night become less frequent, and tree seedlings may

cover actually increased in several of their dry, moist,

be able to survive more frequently (see chapter 14). A

and wet study areas. In addition to the direct effects

contributing factor would be that warmer air holds

of drought, various studies have documented variable

more water vapor, which itself is a greenhouse gas that

plant responses to changes in the duration of snow

enables warmer temperatures at night. Expansion of

cover.22

lodgepole pine into dry meadows since about 1870 in

To learn more about the mechanisms whereby sub

Yellowstone National Park apparently is driven by the

alpine meadows might respond to climate warming,

regional warming trend that began at the end of the

small-scale experiments have been conducted using elec-

Little Ice Age.

Similarly, trees became established in

tric heaters. One study in the Colorado Rockies found

some meadows in the Medicine Bow Mountains during

warming increased the water-use efficiency of mountain

this time, apparently for the same reason.20 Such obser-

big sagebrush, enabling higher rates of photosynthesis

vations suggest that tree expansion is likely to continue

early and late in the growing season.23 The results suggest

as the global climate warms over the coming century.

that mountain big sagebrush could become more abun-

19

Mountain Meadows and Snowglades

dant with warming, partly because it has roots that tap

and biological diversity, are important sources of forage

deep soil water late in the growing season.24 In contrast,

and water, and are sensitive indicators of environmental

the rate of photosynthesis in shrubby cinquefoil did not

change.26 With ongoing climate warming, the expan-

increase with warming. Another heating experiment

sion of trees into some meadows is likely to continue

conducted nearby found no effect of warming on plant

where (1) desiccation does not preclude tree survival,

species composition after four growing seasons, possibly

(2) night-time temperatures enable tree seedling estab-

because the earlier drying after snowmelt canceled the

lishment, (3) competition from herbaceous plants is not

benefits of warming.25 Also, nutrients might have been

excessive, and (4) scouring by winter winds is not an

limiting to plant growth, rather than temperature or

important factor. The temptation to maintain some of

water, as observed in alpine tundra (see chapter 14).

these meadows by tree cutting or burning may increase.

If nutrients are a limiting factor for plant growth,

Snow may melt sooner because of warming trends, but

then the effects of warming may be indirect. For exam-

even low snowfall years are likely to produce enough

ple, warming may stimulate growth only after sufficient

water to saturate meadow soils at the beginning of the

time has passed for warming to promote the decomposi-

growing season. Thus, mountain meadows and tundra

tion of soil organic matter, thereby making more nutri-

will continue to be green for much of the summer, even

ents available at a time when water also is available and

during drought yearsin contrast to lowland grass-

temperature conditions are favorable. Invariably, mul-

lands. Consequently, the demand for livestock access

tiple factors are involved. The effects of environmental

to this forage may increase. Ribbon forests and snow-

change are not always immediately obvious; long-term

glades are likely to shift in their locations with changes

experiments are required.

in the timing of snowmelt, but they will continue to


be a feature of the subalpine landscape as long as suf-

In summary, mountain meadows are a highly valued

ficient snowfall and wind-caused drifting occur during

component of mountain landscapes. They add aesthetic

the winter.

229

Upper Treeline
and Alpine Tundra
Chapter 14

Found at the highest elevations, with views of the val-

many of its inhabitants will be unable to survive in the

leys below, alpine tundra is the land of the pika, rosy-

region? To make predictions, conservation biologists are

finch, wind-swept trees, and small plants with big

thinking broadly about the tundra environment and

flowers. During the last Ice Age, the treeline was lower

the ecology of the plants and animals that live there.

and tundra extended over a larger area (see chapter 2).

Subalpine forests give way to alpine tundra at eleva-

Alpine animals could move easily from one mountain-

tions ranging from about 11,480 feet in the Medicine

top to another. Since then, as the climate has warmed,

Bow Mountains, in southern Wyoming, to about 9,840

the tundra has been invaded by trees. Will the tun-

feet in the Beartooth and Bighorn mountains on the

dra habitat be further diminished to the point where

border with Montana (figs. 14.1 and 14.2).1 Slope expoFig. 14.1. Alpine tundra in the
Snowy Range of the Medicine
Bow Mountains. The bedrock
of this glaciated valley is
Medicine Bow quartzite.
Elevation 10,825 feet.

230

Upper Treeline and Alpine Tundra


Fig. 14.2. Upper treeline in the Bighorn
Mountains. The trees are Engelmann
spruce and subalpine fir. Common
plants in the alpine turf are alpine
avens, alpine bistort, alpine forget-menot, alpine sagewort, blackroot sedge,
selaginella, stonecrop, and spike tri
setum. Elevation 9,600 feet.

sure also is important in determining the limits of tree

itous circumstances. William K. Smith and his associ-

growth, with treelines often at higher elevations on

ates, working in the Medicine Bow Mountains, found

the warmer south slopes with less snow cover. North

that seedling establishment was facilitated when the

and west slopes tend to have more snow later into the

area around the seedling favored a moderate amount of

spring, which causes the treeline to be lower.2

snow cover, providing protection from abrasion and desiccation during winter. As important, they found that

Krummholz

seedlings required at least a minimum of plant cover to


provide shade during the summer, not just from intense

The trees commonly found at treeline are Engelmann

alpine sunlight during the day but also from the cold

spruce, subalpine fir, limber pine, and, in the Greater

sky at night.4 Because of the cold-sky effect at night,

Yellowstone Area, whitebark pine.3 All are short at this

unshaded leaves could be colder than the minimum air

elevation. On exposed sites the trees have a windswept,

temperatures on near-freezing summer nights, which

shrubby growth form known as krummholz, a German

was found to inhibit photosynthesis the following day.

word meaning twisted wood. Some have vertical trunks

Summertime freezing is detrimental to plants and is

flagged with branches on the leeward side, caused by the

much more likely to occur when they are exposed to

abrasion of blowing ice and snow that prevent branch

the night sky. Moreover, unshaded leaves were consid-

growth upwind (fig. 14.3). Branches and the needle-

erably warmer than the maximum air temperature on

shaped leaves near the ground are denser, forming a mat

a summer day, which greatly increased rates of water

that accumulates a protective snowdrift in winter. The

loss and the likelihood of wilting. With adequate but

shrubby trees are widely spaced in a matrix of alpine

not excessive shading, seedlings have photosynthetic

tundra and subalpine meadows. In some wind-exposed

rates that are sufficient for plant growth and the for-

places they occur in ribbons because of the interactive

mation of mycorrhizaewhich in turn lessens water

effect of wind, snow deposition, and slight changes in

stress. The seedlings future is secured when the young

topography (see chapter 13).

tree is large enough to cause snow accumulation over

The stunted trees can be very old. Most likely they

its own branches for a sufficiently long time during the

became established when a seedling managed to survive

winter, or if the neighboring plants or microtopography

for several years because of a rare combination of fortu-

cause the needed snow drifting. This research suggests

231

232 Foothills and Mountains

Fig. 14.3.Trees at upper treeline often have the krummholz


growth form, with dense lower branches that are protected
from the wind by snow during the winter. Upper branches
develop only on the leeward side of the trunk because of
abrasion from blowing ice and snow. Most krummholz trees
are Engelmann spruce and subalpine fir, though whitebark

pine can have a similar growth form in northwestern


Wyoming. Common plants in the adjacent alpine fellfield
include alpine avens, alpine bistort, alpine sagewort, and spike
trisetum. Photo taken on Libby Flats in the Medicine Bow
Mountains. Elevation 10,800 feet.

that understanding the elevation of treeline depends as

matter and nitrogen under the trees. For a time, some

much on where tree seedlings are able to become estab-

alpine plants may do better on soils left behind after

lished as on climatic factors that limit growth.5

the trees have passed. However, wind scouring of the

Often the shrubby tree islands are initiated in slight

organic-rich soil most likely resumes after the protec-

depressions, where water is more readily available and

tion of windward branches is lost, leaving the tundra

seedling establishment is more likely to occur. Eventu-

with a soil similar to that prior to the establishment of

ally, after many years, the lower branches of the saplings

the first tree seedling.

are pressed to the ground by heavy snow and begin to

The islands of krummholz influence plant distribu-

produce roots. This process, known as layering, contrib-

tion in other ways, primarily through drifting snow

utes to the horizontal dimensions of the krummholz

and snowmelt water. 8 The snow provides protection

mat, creating the impression many years later of sev-

during the winter, and the temperature under snow

eral trees growing together.6 In fact, most tree islands

is warmer than on the soil surface of snow-free tun-

are one or a few clones developed from one or a few

dra. Downslope from the drifted snow, water is more

seedlings.

readily available during what would often be a dry

Once established, the patches of krummholz have

growing season, contributing to the amount of plant

several interesting features. First, they move very slowly

growth that can occur there. Commonly found under

downwind, perhaps an inch or so each yearthe result

the shrubby trees are such plants as dwarf huckleberry,

of abrasion on the upwind side and layering on the

mountain gooseberry, and other species associated

downwind side.7 As they move, the tundra soil under

with the less severe environment of subalpine forests

the krummholz is changed, with more soil organic

and meadows.

Upper Treeline and Alpine Tundra

Though wind surely contributes to treeline forma-

violet radiation and the excessive reflection of solar

tion through its effect on snow accumulation and new

radiation from snow onto leavesboth of which can

seedling establishment, there is compelling evidence

retard physiological processes.

that heat deficiency is another cause.9 By their nature,


trees have woody stems and branches that must be
maintained by carbon-rich compounds produced by

Avalanches

photosynthesisa temperature-dependent process. As

Snow avalanche tracks are a conspicuous feature of

elevation increases, temperatures become too low for too

treelines where large volumes of drifting snow accumu-

much of the growing season to allow the amount of pho-

late, such as on the lee side of high ridges and where

tosynthesis required for the year-round maintenance of

snow slides are common (fig. 14.4). Such areas are par-

large plants. Hence, the trees become shorter and even

ticularly common in Colorado, northwestern Wyo-

shrubby in nature, and at some point, small seedlings

ming, and parts of the Northern Rocky Mountains.

cannot survive either. Worldwide, trees are absent where

Usually the snow accumulates in the alpine zone and

the growing season is shorter than 3 months and the


growing season average temperature is less than about
43F.10 Perennial herbaceous plants and small shrubs
are clearly better adapted for the cold alpine environment, probably because less photosynthesis is required
to maintain small plants and the roots and rhizomes of
such plants store carbohydrates for the production of
new leaves early the following year. Small plants also are
more likely to be protected by snow cover during winter.
Predictably, some trees have adaptations that
enable them to tolerate the cold treeline environment.
For example, treeline conifers have clustered needleleaf arrangements that tend to maintain daytime leaf
temperatures that are considerably warmer than the
ambient air.11 This temperature difference facilitates
photo
synthesis in the cool alpine environment and
occurs because leaf clusters (1) present a more dense
mass for the absorption of solar radiation and (2) minimize wind movement through the needles, thereby
reducing cooling. Limber pine and whitebark pine are
adapted in another way, namely, flexible branches that
reduce breakage from strong winds.12
The challenge of understanding treelines is illustrated by the hypothesis that trees cannot survive in
alpine environments because the summer is too short,
cool, and dry to allow development of the protective
tissues required to tolerate abrasion the following winter. Although plausible, Julian Hadley and William K.
Smith concluded that this hypothesis does not apply
to the Medicine Bow Mountains.13 Their results suggest
that wind-caused abrasion during the winter was more
important than a short growing season. Still other factors affecting treeline elevation are more intense ultra-

Fig. 14.4. Avalanche track in Cascade Canyon, Grand Teton


National Park. Tree damage is less frequent at the lower end
of the track, enabling the trees to grow taller there. However,
as the trees grow larger, they offer more resistance to the avalanche and are more likely to break or be uprooted. The trees
are subalpine fir and Engelmann spruce. The more flexible
shrubs higher on the slope include bearberry honeysuckle,
mountain ash, Rocky Mountain maple, rusty menziesia, and
serviceberry. Elevation at the bottom is 7,216 feet.

233

234 Foothills and Mountains

periodically slides along a topographic depression that


extends into the subalpine forest, ending in the runout
zone. Depending on the amount of snow and other conditions, avalanches can be gentle slides over short distances or thunderous events that break large trees in the
valley bottom. Sometimes several avalanches occur on
the same track in a single winter.
The vegetation of avalanche tracks is different from
the forest on either side or the alpine tundra above. In
the tundra, the track typically has sparse plant cover,
because the catchment area commonly has snow that
persists until midsummer, thereby shortening the growing season. Damage to small alpine plants is minimal,
because snow movement occurs within the snowpack
rather than at the soil-snow interface. Most physical
damage to plants occurs below the treeline, where tall,
woody plants are common. Even there small trees and
flexible shrubs are not damaged. However, trees with
diameters greater than 4 inches usually break, because

Fig. 14.5. White pine blister rust on a whitebark pine branch.


This exotic fungal parasite has killed a large portion of the
whitebark and limber pines in the Rocky Mountains and
is continuing to spread. The infection is fatal when the
parasite grows into the bark of the tree trunk. Photo by Anna
Schoettle.

boles of that size provide too much resistance to the


sliding snow.14

cut cone-bearing branches. Grizzly bears can switch to

Thus, there is a vertical gradient in the plant growth

other food sources, but the loss of whitebark pine over

forms that is correlated with avalanche frequency. Small

such a large area reduces habitat quality for this already

plants and flexible shrubs dominate at the top, with trees

threatened species.

becoming more common lower on the track because of

Millions of trees have been killed wherever five-

longer periods of uninterrupted growth. The largest trees

needle pines occur in North America, but the disease

occur at the bottom, which is rarely affected. Predictably,

currently is most severe in the Rocky Mountains. Infec-

the most severe disturbance occurs when trees are bro-

tion levels vary considerably, with the highest occurring

ken during large avalanches that extend to the runout

in close proximity to the rusts alternate hosts, which

zone. Such disturbances occur every 50100 years.

include wild currants and gooseberries (Ribes spp.) plus


a few other species.16 Some populations of the trees

White Pine Blister Rust at Treeline

seem resistant to the disease, which gives conservation


biologists hope that future generations of the pines will

A disturbing development is the invasion of an exotic

be more resistant. That seems to have happened around

disease that kills whitebark pinethe common species

the Great Lakes and farther to the east, where eastern

at treeline in the Greater Yellowstone Area and North-

white pine is still common today, even though it was

ern Rocky Mountains.15 Known as white pine blister rust

often infected by the blister rust in the mid-1900s. A

(Cronartium ribicola), this fungal parasite also kills lim-

complicating factor in western states, though, is that

ber pine and bristlecone pine in the Southern Rockies

whitebark and limber pine, even if generally resistant

(fig. 14.5). All three of these pines tolerate severe envi-

to the rust, may nevertheless be sufficiently weakened

ronments, and their peanut-sized seeds historically have

by the disease to be killed by the native mountain pine

provided an important food supply for grizzly bears,

beetle (see chapter 11). Research on this perplexing prob-

red squirrels, Clarks nutcracker, and other animals (see

lem and the resilience of the ecosystem is ongoing.17 As

chapter 10). The bears are able to build up fat reserves

at lower elevations, the nature of the ecosystem may

as they gorge themselves on the pine nuts, which some-

change considerably because of inadvertently intro-

times are accessible on the ground because red squirrels

duced invasive species.

Upper Treeline and Alpine Tundra

Surviving in the Alpine Tundra


Above treeline, many of the shrubs and herbs that survive are also found in the Arctic. Both kinds of tundra
have cool, short growing seasons. However, there are
prominent differences: the Arctic has long summer days
with little diurnal temperature fluctuation, low ultraviolet and visible light intensities, and permafrost over
much of the landscape. In sharp contrast, alpine zones
have high intensities of ultraviolet and visible light,
shorter summer photoperiods, and great diurnal temperature extremes in the summeroften reaching more
than 68F during the day and dropping below freezing
at night. Also, permafrost is rare in the alpine tundra.18
Wind and cold are clearly important aspects of the
alpine environment, but drought is another factor
19

affecting plant survival.

Evapotranspiration is more

rapid because of lower atmospheric pressures at high elevations, which increases the probability of plant water
stress. Also, water uptake by roots is slower when the
soil is cold. Various investigators have noted the similarity of desert and alpine plants in terms of leaf size and
other structural features. Nutrients can be limiting as
well, which is partially compensated for by the presence
of mycorrhizae and nutrient resorption prior to leaf fall.
Notably, most of the root system, which has two to ten
times more biomass than the stems and leaves above
ground, is in the top 4 inches of the soil, where nutrient
availability is highest and the temperature is relatively
warm during summer.
Arguably, the alpine tundra has the most severe
conditions for plant growth to be found anywhere in
the region: extreme temperature fluctuations during a
24-hour period (which is very difficult for most plants
to tolerate); frequent freezing and thawing; rapid rates
of drying (away from melting snowdrifts); nutrient limitations; and a cool, short growing season. Still, several
hundred plant species tolerate this environment in
Wyoming alone.20
The predominant growth forms of alpine plants are
perennial herbs (including forbs, grasses, and sedges)
and short woody or semi-woody shrubs. Most biomass
is belowground. Annuals are rare because seedling
establishment is difficult in the alpine environment.21
Lichens are common, but they are not a dominant feature of alpine vegetation, as they are in the arctic tundra.

Annual plant growth of the tundra has been estimated


at 40 g/m 2/year on dry, windswept sites and 300 g/m2/
year on more protected, relatively moist sites.22 Such
growth rates, although low compared to subalpine forests, are reasonably high, considering that the growing
season is only 3075 days long.
Predictably, some unusual adaptations have evolved
to allow alpine plant growth.23 For example, all alpine
plants are capable of photosynthesis and growth at cold
temperatures, and many are drought tolerant.24 Evergreen and wintergreen leaves are common, enabling
growth early and late in the season, and even while
still covered with up to 20 inches of snow. Root systems
tend to be shallow, probably to facilitate nutrient uptake
as decomposition in the surface soils occurs, and the
roots usually have mycorrhizae that greatly increase the
surface area available for nutrient acquisition. Alpine
plants also seem well adapted for mitigating the adverse
effects of a short growing season. For example, alpine
pennycressa diminutive herbgrows from upper
treeline down into the foothills, but the higher populations complete their growth and reproduction more
rapidly than those at lower elevations.25
Another plant adaptation is the cushion plant
growth form, with branches and leaves that grow in
dense mats close to the soil surface (for example, moss
campion, alpine forget-me-not, and alpine phlox). Soil
temperatures during the day can be surprisingly warm
because of the intense sunlight at high elevations. The
dense branches absorb heat from the soilwarmed by
the sunand they slow the movement of air through
the plant, thereby slowing the rate of cooling. Dark color
and epidermal hairs also help maintain warmer leaf and
bud temperatures in this otherwise cool or even cold
environment. Some insects depend on the more moderate microclimate of the cushion plants.26
Reproduction of alpine plants is largely vegetative,
that is, sprouting from roots and rhizomes, because of
the difficulties associated with seedling establishment
in such a rigorous environment. Nevertheless, many
plants have evolved special adaptations for seed production. Some plants have unusually large and showy
flowers, presumably to attract scarce insect pollinators
during the short warm periods when the insects are
active. Flowers with large petals serve as windbreaks
for insects and are thought to provide a warmer micro

235

236 Foothills and Mountains

most of their livesand under snow for much of the


yearthey avoid bone-chilling winds and the coldest
temperatures. 30

Tundra Mosaic, Frost, and Burrowing Mammals


The plant communities of alpine tundra can be easily
classified into one of the following categories: fellfields,
alpine turf, wet meadows, and willow thickets.31 There
are also snow-beds, talus slopes, and boulder fields,
all of which have little or no plant growth. Lakes and
streams are common, with associated wetland vegetation that includes sedge-dominated wetlands as well as
willow thickets (see chapter 5).
Fig. 14.6.The big flowers of this 6-inch-tall alpine tundra
plant, known as alpine sunflower or old-man-of-the-mountain, attract pollinating insects during the short cool summer.

Alpine tundra is commonly discussed at three spatial scales. At the scale of tens of miles, variation is
caused by differences in local climate. 32 The tundra of
the Southern Rockies is different from that of the Cen-

environment (fig. 14.6).27 Because pollination often

tral and Northern Rockies. In contrast, at the scale of a

does not occur during the short summer, a high per-

few feet, abrupt differences can be caused by the pres-

centage of alpine plants are capable of producing seeds

ence of boulders, which absorb and store considerable

without fertilization, a process known as apomixis.

heat from the intense sunlight at high elevations, and

Animals have equally interesting adaptations. Aside

then release the heat at night to the benefit of plants

from fish, all vertebrates in the alpine zone are warm-

growing nearby. Also, snow accumulates on the lee side

blooded endotherms, which is adaptive as long as they

of boulders, and water from rain and condensation

have the food required to maintain body temperature.

drains rapidly to the soil below the boulders, where it

No alpine amphibians or reptiles have been found.

is less likely to evaporate. Because of these influences,

Some animals migrate up and down the mountain to

boulders provide warmer, wetter microenvironments

find food and shelter. During the winter, a few species


of birds leave the region altogether, like the American
pipit. Other animals have the extraordinary ability
to stop eating while they hibernate, such as the marmot. Mice, voles, and pocket gophers consume roots
and other accessible plant materials in their burrows,
even under the snow, whereas the pika works all summer long to create a cache of food deep in the boulder
fields and talus slopes where it lives (fig. 14.7). Known
as haypiles and composed largely of alpine avens and
other forbs, these caches are quite large, weighing on
average about 60 pounds in the Colorado alpine. For
comparison, each pika weighs less than 7 ounces.28
There can be six to eight pikas per acre, each one feeding outward from its nest. 29 Through their burrowing,
all small mammals contribute to soil development and
have effects on the vegetation where they live. Tons
of soil are aerated. By living in soil or under rocks for

Fig. 14.7.The American pika, 68 inches long, lives in alpine


boulder fields and talus slopes, where it creates haypiles for
warmth and food during the winter. It does not hibernate.
Copyright Wendy Shattil/Bob Rozinski.

Upper Treeline and Alpine Tundra

SNOW COVER
Snowfree

Winter snow
cover

TOPOGRAPHIC
POSITION
Fellfield

Summit
or
ridge

Persistent
snow

Leeward

or tree islands typically have drifted snowbanks that


persist until midsummer. Few if any plants can survive
where snow persists late into summer, but plants growing downslope from late-lying snow have a dependable

Upper
slope

source of water for most of the growing season. Consequently, the vegetation there has higher cover and

Turf

growth. Moreover, the dominant plants of these mead-

Tufted
hairgrass meadow

Snow
accumulation

tion during the winter (fig. 14.8). The lee sides of ridges

Sedge meadow

Lower
slope
Ravine

Willow or fen

Windward

WIND EXPOSURE
Fig. 14.8. Distribution of different kinds of alpine vegetation
in relation to topographic position, wind exposure, and snow
cover. Adapted from Johnson and Billings (1962).

ows are different (table 14.1).


Perhaps the most unique process that causes pattern in the alpine landscape is cryoturbationthe freezing and thawing of moist soils that occurs on time
scales of a single day or a year. Soil water expands as
it freezes, pushing comparatively large objects, such
as stones and boulders, along the path of least resistanceusually toward the surface. After thousands of
freeze-thaw cycles, the larger objects are sorted from
smaller ones, creating patterned ground, such as frost
boils, stone nets, and stone polygons (fig. 14.9). On a

that allow some plants to grow where they might not

24-hour cycle, needle-ice forms and melts, which can

otherwise.

lead to the formation of frost boils on the soil surface

At an intermediate scale, differences are caused by

that damages the roots of seedlings (fig. 14.10). 33

topography and the presence or absence of tree islands,

Another frost-related process is solifluction, in which

as discussed previously. On windswept ridges or flats

frost and gravity combine to move wet soil downslope.

are fellfields, dominated by cushion plants and other

Solifluction occurs most often on lee slopes, where

drought-tolerant plants that require little snow protec-

snow accumulates and the soil becomes wet, gradually

Fig. 14.9. Alternating freezing and


thawing over long periods causes large
stones and rocks to be separated from
fine-textured soil particles, forming
stone polygons or nets, such as in this
area in the Beartooth Mountains. The
alpine turf is dominated by various
alpine sedges along with alpine avens,
alpine bistort, alpine sagewort, snow
willow, tufted phlox, and other species.
Elevation 10,400 feet.

237

Table 14.1. Some characteristic plants of alpine plant communities in Wyominga

Common name

Latin name

Fellfield

Alpine turf

Wet
meadow

Willow
thicket

FORBS
Arctic alpine forget-me-not

Eritrichium nanum

Alpine leafybract aster

Symphyotrichum foliaceum

Alpine mountain sorrel

Oxyria digyna

Alpine sagebrush

Artemisia scopulorum

American bistort

Polygonum bistortoides

Creeping sibbaldia

Sibbaldia procumbens

Dwarf clover

Trifolium nanum

Elephanthead lousewort

Pedicularis groenlandica

Graylocks four-nerve daisy

Tetraneuris grandiflora

Hookers mountain avens

Dryas octopetala var. hookeriana

Lesser spikemoss

Selaginella densa

Moss campion

Silene acaulis

Parrys clover

Trifolium parryi

Parrys lousewort

Pedicularis parryi

Redtop stonecrop

Rhodiola rhodantha

Rocky Mountain nailwort

Paronychia pulvinata

Rocky Mountian pussytoes

Antennaria media

Ross avens

Geum rossii

Sticky polemonium

Polemonium viscosum

Tufted phlox

Phlox pulvinata

Varileaf cinquefoil

Potentilla diversifolia

White marsh marigold

Caltha leptosepala

Western yarrow

Achillea millefolium var. occidentalis

GRASSES
Alpine bluegrass

Poa alpina

Timberline bluegrass

Poa glauca ssp. rupicola

Alpine timothy

Phleum alpinum

Purple reedgrass

Calamagrostis purpurascens

Rocky Mountain fescue

Festuca saximontana

Spreading wheatgrass

Elymus scribneri

Spike trisetum

Trisetum spicatum

Tufted hairgrass

Deschampsia caespitosa

Bellardi sedge

Kobresia myosuroides

Blackroot sedge

Carex elynoides

SEDGES

Curly (rock) sedge

Carex rupestris

Drummonds sedge

Salix drummondiana

Upper Treeline and Alpine Tundra

Common name

Latin name

Fellfield

Alpine turf

Wet
meadow

Willow
thicket

Ebony sedge

Carex ebenea

Obtuse sedge

Carex obtusata

Water sedge

Carex aquatilis

Alpine laurel

Kalmia microphylla

SHRUBS
Alpine willow

Salix petrophila

Cascade willow

Salix cascadensis

Diamondleaf willow

Salix planifolia

Grayleaf willow

Salix glauca

Snow willow

Salix reticulata

Resin birch

Betula glandulosa

A dash indicates that the plant is absent or uncommon.

slipping or creeping downward. Formed over long peri-

about whether a stable tundra ecosystem ever devel-

ods, solifluction lobes or terraces cause a distinctive pat-

ops. However, gradual changes occur over time scales

tern (fig. 14.11).

of centuries. To illustrate, with gradual weathering and

Cryoturbation and solifluction, combined with the

soil development, boulder fields develop into fellfields,

burrowing of pocket gophers (fig. 14.12), are thought

which can become dry meadows, commonly known as

to be continual sources of disturbance to alpine vegeta-

alpine turf. Similarly, wet meadows dominated by water

tion, creating a mosaic of patches and raising questions

sedge develop into somewhat drier meadows dominated


by tufted hairgrass. On a smaller scale, episodes of cryoturbation cause changes in the dominant plants. 34
Disturbances that occur in alpine tundra are distinctly different from those occurring in ecosystems at
lower elevations. However, numerous plants and animals have evolved adaptations for this environment
and are not found elsewhere. Some observers refer to
tundra as fragile, because recovery after disturbances is
slow. To be sure, excess livestock grazing and damage
caused by foot traffic and off-road vehicles can cause
undesirable changes that persist for many decades. 35

Nitrogen Deposition
Nitrogen fertilization experiments in alpine tundra
Fig. 14.10. Frost boils, such as this one on Libby Flats in the
Medicine Bow Mountains, create small patches in the tundra
vegetation. Burrowing mammals, like the Wyoming ground
squirrel, may be beneficiaries of the loose soil, or they may be
involved with forming such features. Elevation 10,500 feet.

commonly lead to an increase in plant growth, indicating that nutrients can be limiting along with the cool,
short growing season.36 Nitrogen-fixing plants do occur,
such as alpine clover, but they are not widespread.

239

240 Foothills and Mountains


Fig. 14.11. Solifluction lobes or terraces
develop over a fellfield on slopes in
alpine tundra, such as in this area in
the Beartooth Mountains. Solifluction
creates different environments for plant
growth over short distances. Water
seeping through the lobes provides an
ideal environment for barrenground
willow. Elevation about 10,600 feet. See
Johnson and Billings (1962).

Limiting nutrients tend to be conserved in plant and

nutrient source. One study concluded that the eutro-

microbial biomass and will not normally be leached or

phication of Rocky Mountain aquatic ecosystems from

lost from the soil except after disturbances.37

atmospheric pollutants is analogous to the acidification

It was unusual, then, when investigators observed


high concentrations of nitrogen in the creeks drain-

of terrestrial ecosystems associated with the same pollutants in northeastern North America.42

ing the alpine zone of Rocky Mountain National Park


and other parts of the Colorado Front Range. 38 No obvious disturbances were evident, but further research led
to the conclusion that the addition of nitrogen in air
pollutants, mainly nitrous oxides, was causing a condition known as nitrogen saturation, when the amount
of nitrogen added to the ecosystem is essentially the
same as the amount lost.39 Nutrient enrichment, known
as eutrophication, leads to changes on the upland and,
downstream, enables algal blooms that can reduce fish
productivity and recreational opportunities.40
Notably, the added nitrogen apparently is first
deposited at high elevations in the tundra.41 Referred
to as nitrogen hotspots, such areas have large amounts
of bare rock and a small amount of biomass for nitrogen uptake. Moreover, much of the nitrogen is added in
snowfall that is blown into drifts, thereby concentrating the nitrogen in places that have the least amount of
plant growth and storage capacity. In contrast, the subalpine forest and many subalpine meadows have higher
nitrogen-storage capacities and currently are less of a

Fig. 14.12. Pocket gophers, though rarely seen, are common


in alpine tundra and subalpine meadows and often burrow in
the soil under snowdrifts, where they eat roots. The soil from
their burrows is pushed into tunnels through the snow, which
become middens when the snow melts. These small mammals
aerate the soil and facilitate nutrient cycling. The plants in
this area are sparse, because snow cover persists until early
July.

Upper Treeline and Alpine Tundra

The response of tundra plants to nitrogen enrich-

higher precipitation (see chapter 3). Tree establishment

ment is currently under investigation. University of

could be favored by both warmer temperatures and less

Colorado ecologist Timothy Seastedt and his associ-

drought stress, but apparently such conditions have

ates have observed that willows are expanding their

not yet developed in some areas, such as in northern

distribution in the alpine zone, possibly because of

Colorado.49 Moreover, with ongoing disturbances (such

the combination of added nitrogen and climate warm-

as frost heaving, burrowing by small mammals, wind

ing. This result might be expected, he suggests, in wet

abrasion, and herbivory), the chances of new tree seed-

areas that are protected from wind scour, that is, places

ling establishment in the tundra are still lowand prob-

where organic matter accumulates as well as nitrogen-

ably will be for many years to come.50 Alpine ecologist

enriched snow and water. The more exposed dry mead-

Christian Krner concluded in 2012 that the advance

ows and fellfields would receive less nitrogen and would

of treelines will always lag behind climate change by

be less likely to change. Thats the hypothesis.

at least 50100 years, considering how long it takes for

43

new seedlings to become established and grow to tree

Advancing Treelines and Climate Change

height in the alpine environmenteven if it is slightly


warmer.

It is well known that treelines have moved up and

But not all alpine treelines are the same. For exam-

down during the past 10,000 years (see chapter 2).44 But,

ple, new tree seedlings have been found above treeline

if snowmelt is occurring earlier in the spring because

on relatively moist sites in the Front Range of Colo-

of climate warming, and glaciers are retreating, is the

rado.51 Also, in the Uinta Mountains of Utah, photos

alpine treeline moving upward throughout the region?

taken of the same place in 1870 and 2000 indicate a rise

Answering this question has been the subject of con-

in treeline of 180550 feet.52 In the same area, plants

siderable research, especially since the mid-1990s.45 The

other than trees have been observed moving upward in

results are mixed and offer an interesting example of

elevation through seedling establishment. Such upward

the complexities of interpreting the effects of climate

migration might have occurred even without the cur-

change. In general, the same results should not be

rent warming trends, but vegetation response to warm-

expected for alpine treelines everywhere.

ing clearly varies considerably.

Fundamental to treeline advance is the ability for

As pertinent to the effect of warming on alpine plant

new tree seedlings to become established higher on the

distribution is the paradox that earlier snowmelt in the

mountain than their parent trees could, that is, higher

spring from warming increases the probability of frost-

than the krummholz that already exists at treeline.46 In

related plant death or damage. Some alpine plants are

some places, newly established tree seedlings in the tun-

sensitive to cold temperatures after they begin growth,

dra have been hard to find, suggesting that trees at this

which is delayed by snowcover. If the snow melts ear-

elevation are not a good indicator of climate change.47

lier, frost damage is more likely.53 Global warming that

How could this be? One study in Rocky Mountain

causes higher mortality could retard the rate at which

National Park concluded that the treelines are relictual

treeline plants respond. Also, earlier snow-free dates

they persist because adult trees tolerate the present-day

could lead to colder soils and more cryoturbationtwo

climate, even though the more delicate seedlings can-

factors that reduce the chance of seedling survival.

not.48 In other words, the tree seedlings that founded

Climate warming also will affect alpine animals that

the tree islands of today became established at a time

are restricted to mountain tops. On mountains with less

when the climate was more favorable. Numerous seed-

land in the alpine zone, the populations of alpine species

lings were found invading openings in the subalpine

are smaller; and if treelines advance upward because of

forest below, especially in moist areas, but hardly any

climate change, their populations will become smaller

were found in the drier openings at treeline or above.

still, increasing the chances of local extinction. Ani-

In addition to warmer temperatures, one of the

mals on one mountain may be able to travel to others

predictions for climate change in the alpine zone is

nearby if they can fly or if they are able to survive or

241

242 Foothills and Mountains

tolerate the subalpine forest ecosystem as they move

able to survive below treeline where subalpine meadows

from one alpine zone to another. Populations of the

and talus slopes are in close proximity.

same species farther north, where there is more tundra


because treelines are lower, may have a greater chance

Overall, sorting out the causes of shifting treelines

of surviving.

and nutrient enrichment is a challenge that requires

Such discussions invariably consider the plight of

an understanding of microclimate; wind; soil; cryo

the American pika, a small mammal in the rabbit fam-

turbation; the effect of cold night skies; the chemistry of

ily that essentially is restricted to alpine tundra (see

the atmosphere and soil; the various factors that influ-

fig. 14.7). It nests in the crevices of boulder fields and

ence streamwater chemistry; and the adaptations of the

frequently is seen scurrying from one rock to another,

plants, animals, and microorganisms found on land

often stopping to bark along the way. If the forest mar-

and in water. Treelines are likely to expand upward,

gin continues to move upward because of global warm-

slowly at first. The alpine tundra, though far removed

ing, pika habitat will be squeezed into smaller and

from urban-industrial developments at low elevations,

smaller patches. At present, the alpine treeline appears

is nevertheless another bellwether of changes under

to be moving upward very slowly, if at all, and pikas are

way throughout the region.

Part Five Landscapes of

Special Interest

This page intentionally left blank

The Greater
Yellowstone Ecosystem
Chapter 15

The Greater Yellowstone Ecosystem (GYE) covers some

depth of 3 inches as far away as present-day Iowa. This

40,000 square miles and encompasses the largest wild-

time, with the loss of so much material within a mat-

land in the contiguous United States. It includes two

ter of hours or days, the center of Yellowstone country

national parks, seven national forests, ten designated

collapsed on itself, forming a giant depression known

wilderness areas, three wildlife refuges, and a variety

by geologists as a caldera. The underlying magma was

of other public, tribal, and private lands (fig. 15.1).1

still hot, however, and a long series of lava flows began

The influences of modern industrial society have been

filling in the depression, with major flows occurring as

minorcompared to those on the surrounding area

recently as 70,000 years ago. The thermal features that

and with the reintroduction of the gray wolf in 1995,

so many visitors enjoy today are evidence that Yellow-

all of the GYEs original pre-Columbian mammals are

stone is still underlain by a hot spot. The Yellowstone

present. Reading a list of the names of Yellowstones

caldera occupies about a third of YNP. The high cliffs

largest mammalselk, bison, wolves, moose, and griz-

along the north side of the Madison River are part of the

zly bearsis akin to contemplating Tyranosaurus rex,

north rim, Mount Washburn (10,223 feet) lies on the

Apatosaurus, and Pterodactylus coexisting in a landscape

northeastern edge, Mount Sheridan (10,298 feet) is on

with mysterious plumes of steam and water emanating

the southern edge, and Canyon Village is in the bottom

from the earth. Little wonder that the early explorers

at 7,915 feet. West Thumb and the central and northern

referred to the area as a wonderland.

portions of Yellowstone Lake also lie in the caldera.

In the Eocene, 55 to 35 million years ago, a series of

The processes that formed the Teton landscape, just

volcanic eruptions literally buried the mountain ranges

south of YNP, are different but no less amazing.5 Some

in and around what is now Yellowstone National Park

40 million years after mountain formation to the north,

(YNP) (see chapter 2). New forests developed after each

east, and south, the TetonJackson Hole region was still

eruption, one on top of the other. Fossil tree trunks are

relatively flat. The ancient Precambrian rocks were bur-

still visible today (fig. 15.2).3 The Absaroka Range, on

ied under nearly 4 miles of younger sedimentary and vol-

the horizon for someone looking east across Yellow-

canic rocks. Then, about 9 million years ago, uplifting

stone Lake, is a remnant of Eocene volcanic activity.

began along the 50-mile-long Teton fault, creating the

Most of the Eocene deposits to the west of the Absaro-

youngest range in the Rocky Mountains. As the Tetons

kas are gone, literally blown away by some of the most

rose, the softer rocks eroded. Today, essentially all sedi-

dramatic geological events in Earths history.4 Volcanic

mentary strata have been stripped from the steep eastern

activity occurred again, about 650,000 years ago, with

slopes and the highest peaksexcept for a telltale sand-

the ash deposited across much of North Americato a

stone remnant on the top of Mount Moran. Geologists

245

246 Landscapes of Special Interest

Fig. 15.1.Elevation map for the Greater


Yellowstone Ecosystem. See fig 1.3 for a
map showing land under the jurisdiction
of different agencies, and fig. 1.5 for the distribution of different kinds of vegetation in
the area. The land area covered by this map
is about 50,000 square miles. Cartography
by Ken Driese.

know this remnant is part of the same Flathead sand-

the bottom of Jackson Hole to the top of Grand Teton (fig.

stone that now lies about 6 miles below the valley floor.

15.3). As expected for a mountain range created along a

Slippage along the fault continues at about 12 inches

fault line on the east side, the west slopes are gentler and

per century in some places, with the valley floor sink-

are still covered by sedimentary strata (fig. 15.4).

ing while the mountains rise. Seven peaks in the Tetons

With regional uplifting, continental cooling, and cli-

rise above 12,000 feet, with Grand Teton the highest

mate change, glaciers crept over the Yellowstone region

at 13,770 feet. Nowhere else in the Rocky Mountains is

and then retreated numerous times (see chapter2). The

there such an abrupt rise in elevation6,500 feet from

most recent glaciation, known as the Pinedale, began

Fig. 15.2. (left) Petrified tree trunks on a ridge south of the


Lamar River valley, the remnants of forests buried by volcanic
deposits in what is now Yellowstone National Park during
Eocene eruptions, about 50 million years ago. The fossil trees
in the region include specimens of chestnut, magnolia, maple,
oak, redwood, sycamore, and walnutnone of which grow in
the area now. At the present time, Douglas-fir is common on
this steep north-facing slope. Elevation 8,400 feet.
Fig. 15.3. (below) The Tetons rise abruptly above the Snake
River and Jackson Hole to a maximum elevation on Grand
Teton of 13,770 feet. Shrublands dominated by mountain
big sagebrush are characteristic of the flat glacial outwash
plains, at an elevation of 6,600 feet, including the lower
terrace carved during the Pleistocene Epoch. Blue spruce,
Engelmann spruce, narrowleaf cottonwood, river birch, silver
buffaloberry, and various willows are common in the riparian
woodlands. Douglas-fir occurs on the slopes above the river.
See also fig. 2.6.

248 Landscapes of Special Interest


Teton Mountains

WEST

EAST

Grand Teton
13,770 ft.

n
to t
Te aul
F
Precambrian

sea level
0
0

Jackson Hole

Blacktail Butte

Jackson

Quaternary

Mesozoic

Paleozoic

10 km
10 mi

Fig. 15.4. Geologic structure of the Teton Range and Jackson


Hole. Continuing displacement along the Teton fault has
created a mountain range without foothills. The Quaternary

deposits are mostly alluvium, often referred to as glacial outwash plains, deposited as the glaciers melted. Adapted from
Love et al. (2003).

about 80,000 years ago and reached its maximum extent

in pothole topography. Also known as kettles, these

some 25,000 years ago.7 At its peak, ice covered almost all

potholes were created by the melting of huge blocks

of what is now YNP and was about 4,000 feet thick above

of ice buried in the moraines. Water-filled depressions

the Yellowstone Lake basin. Ice buried and subsequently

remained after the ice disappeared (see chapter 5).

carved many mountains in the area, including Mount

Based on fossil pollen records from lake sediments,

Washburn and Mount Sheridan. The ice also dammed

paleoecologist Cathy Whitlock and colleagues deter-

the Yellowstone River, causing Yellowstone Lake to

mined that tundra was prevalent in much of the GYE

expand and cover the area that today is Hayden Valley.

about 14,000 years ago.9 By about 11,500 years ago,

Fine-textured sediments deposited during that time in

Engelmann spruce began to dot the tundra where a

the bottom of the lake now support extensive meadows.

more fertile substrate had developed from andesitic

Several thousand feet of ice also covered the north-

rocks. Whitebark pine and subalpine fir joined the

ern part of Jackson Hole at times,8 and subsequent gla-

spruce about 11,000 to 9,500 years ago, but the less fer-

ciers flowing down the canyons of the Tetons created

tile rhyolitic soils of the Yellowstone Plateau remained

terminal moraines at the foot of the mountains, form-

unforested during this period. The climate gradually

ing lakes known today by the names of early explor-

became warmer, though it was still cooler and more

ersJackson, Bradley, Jenny, Leigh, Phelps, and Taggart.

humid than it is today. Continued warming led to the

Glaciers also shaped the jagged peaks and U-shaped val-

expansion of lodgepole pine throughout the region.

leys of the Teton, Beartooth, and Wind River Mountains

Whitlocks data indicate that limber pine was present at

(see fig. 2.6).

low elevations and that Douglas-fir expanded to higher

The glaciers began to shrink as the climate warmed,

elevations about 9,500 to 4,500 years ago. Subsequently,

about 20,000 years ago, leading to the gradual elimina-

the climate became cooler and wetter at high eleva-

tion of the vast Yellowstone ice field. Evidence suggests

tions, enabling spruce and fir to become more common,

that a glacier persisted in Jackson Hole until as recently

along with lodgepole pine, as they are today. Douglas-fir

as 9,000 years ago. It would have been witnessed by

now grows at lower elevations.

some of the earliest Native Americans as they entered

Today, the GYE spreads over fourteen mountain

the valley. With rapid thawing of huge masses of ice,

ranges and is a major source of water for three major

rivers became swollen with water and sediment. Great

river systemsthe Green flowing into the Colorado, the

floods occurred as moraines and ice dams collapsed,

Snake into the Columbia, and the Yellowstone into the

depositing large outwash plains (such as Antelope Flats)

Missouri and Mississippi.10 About a third of the land is

and scouring stream bottoms. The extensive terraces

forested, with lodgepole pine, Engelmann spruce and

along the Snake River were formed during this time (see

subalpine fir, and aspen dominating an estimated 20,

fig. 15.3). Further evidence of the glaciers is apparent

12, and 3 percent of the land, respectively. About 10

The Greater Yellowstone Ecosystem

percent is alpine tundra. Shrublands dominated by big

toward Yellowstone and the Tetons.17 He was followed

sagebrush are characteristic of the lowlands, growing on

by various trappers, including Osborne Russell, who

about 25 percent of the GYE.

wrote about the environment, wildlife, and people of


the Yellowstone region in the early 1800s.18
Today YNP occupies an area of 3,468 square miles.

Yellowstone National Park

The highest mountain peaks occur in the Absaroka

Early explorers along the headwaters of the Missouri

Mountains on the eastern boundary, rising to slightly

and Columbia rivers returned with tales of abundant

more than 10,900 feet; the lowest elevation is at Gar-

fish and wildlife, boiling mud, steam rising from holes

diner in the northwestern corner, where the Yellow-

in the earth, hot springs, spectacular waterfalls, and a

stone River leaves the park at 5,314 feet. More than 150

deep canyon with walls of yellow stone. So incredible

lakes cover 5 percent of the park, all small except for

were their stories that few took them seriously. Some

Yellowstone, Shoshone, Lewis, and Heart lakes. Soils

newspapers hesitated to print their reports. In 1870,

have developed mostly on two kinds of volcanic rock

General Henry D. Washburn led an expedition to eval-

the infertile rhyolite and more fertile andesiteplus

uate some of the claims. Cornelius Hedges, a member

the gravel, sand, silt, and clay deposited by retreating

of the expedition, was probably trying to preserve his

glaciers. The highest mean annual precipitation, up to

reputation as a keen observer when he commented,

70 inches, occurs on the high Madison and Pitchstone

I think a more confirmed set of skeptics never went

plateaus in the southwestern corner of the park and in

out into the wilderness than those who composed our

the Gallatin Mountains in the northwest, with the inte-

party, and never was a party more completely surprised

rior and lower elevations of the park receiving much

and captivated with the wonders of nature.12 Hedges

less (only about 10 inches at Gardiner, Montana). Two

complained of the wind and cold in August but, after

distinct climate regimes are recognized in YNP and the

11

walking along the rim of the Grand Canyon of the Yel-

GYE. In the areas located generally south and west of

lowstone River, he was filled with too much and too

the Continental Divide, winters are wet and snowy, and

great a satisfaction to relate.13

summers are relatively dry; but in areas located gener-

A year later, in 1871, Ferdinand V. Hayden led a sci-

ally north and east of the Continental Divide, winters

entific expedition to the region, sponsored by the U.S.

are somewhat drier, and summers are relatively wet.

Congress. Haydens report, with stunning photographs

Some vegetation patterns reflect this climatic pattern;

by William H. Jackson and paintings by Thomas Moran,

for example, aspen is more abundant in the winter-wet

was instrumental in Yellowstone being established in

region in the southern part of the GYE than in the sum-

1872 as the worlds first national park.

mer-dry region to the north.19

14

The accolades

continued. In 1875 General W. E. Strong wrote, Grand,

Rivers are a popular feature of the Yellowstone land-

glorious, and magnificent was the scene as we looked

scape, because they are big and clean. Pulses of sedi-

upon it from [Mount] Washburns summit. No pen can

ment-rich water occur following natural erosion events,

write itno language describe it. Today the wonders

especially in the spring and after heavy thunderstorms

of Yellowstone still entice and excite several million

over recently burned forests, but most of the time the

visitors each year.

rivers run clear. They also offer some of the finest trout

15

Of course, the ancestors of modern tribes had been

fishing in the world, with the native cutthroat trout con-

present 8,00010,000 years before these explorers

sidered a prize. Canyons, point bars, cut banks, terraces,

arrived.16 Aside from hunting and gathering their food,

and broad floodplains are part of the riparian mosaic

they collected and traded obsidian, a hard volcanic rock

(see chapter 4). Rapids and waterfalls are common. The

highly valued for spear points, arrowheads, and cutting

Grand Canyon of the Yellowstone River is about 1,500

implements. When Europeans first arrived, the major

feet deep, passing through rhyolite softened by hydro-

tribes living in the area were the Blackfoot, Crow, and

thermal activity. The canyon begins at Lower Falls, with

Bannock-Shoshone. In 1806 John Colter left the Lewis

a drop of about 300 feet, and continues for about 18

and Clark expedition in Montana and traveled south

miles to the mouth of the Lamar River.

249

250 Landscapes of Special Interest

Vegetation of Yellowstone National Park


About 80 percent of YNP is covered by forest (see
fig. 1.5). The parks forests appear uniform to many
visitors, yet ecologists have identified more than
thirty different forest environments (habitat types
and phases). 20 Lodgepole pine predominates over large
areas, especially on the rhyolite plateaus (see chapter
10). Lodgepole is probably the only tree capable of
thriving on such infertile soils; indeed, without lodgepole pine, there might be little or no forest over much
of the park. Some twenty-four habitat types in which
lodgepole pine is a major constituent have been identified. Some of the distinguishing understory species
include globe huckleberry and bluejoint reedgrass on
relatively moist and fertile sites, with grouse whortleberry, pinegrass, and elk sedge on dry and infertile
sites. Engelmann spruce and subalpine fir also are common and in some places dominant, especially on more
fertile substrates derived from andesite or sedimentary
rocks. Whitebark pine is common in high-elevation
forests (fig. 15.5), and Douglas-fir is common in the
foothills (fig. 15.6). 21 Tree-dominated riparian woodlands are not common in the park, but where they do
occur, the principal trees are aspen, balsam poplar,
narrowleaf cottonwood, and sometimes Douglas-fir,
Engelmann spruce, and lodgepole pine. 22

Fig. 15.5. Whitebark pine, shown here at the top of Mount


Sheridan in Yellowstone National Park, is the predominant
treeline species in northwestern Wyoming and the Northern
Rocky Mountains. In many areas the trees have been killed
during the past two decades by native mountain pine beetles
and the non-native white pine blister rust. Elevation 10,308
feet at the summit.

Nonforested land occupies most of the remaining


20 percent of YNP.23 Foothill grasslands dominated

grass. Interspersed throughout the subalpine forests are

by bluebunch wheatgrass and Sandberg bluegrass are

meadows dominated by tufted hairgrass and a variety

found at low elevations near Gardiner, where annual

of sedges and forbs.

precipitation is only about 10 inches. Also found in this

Several species of sedge are especially common in

area are desert shrublands dominated by fringed sage,

fens and wet meadows (see chapters 4 and 5), includ-

Gardner saltbush, greasewood, pricklypear cactus, and

ing beaked sedge, Buxbaums sedge, livid sedge, mud

winterfat. Meadows at low elevations along the Lamar

sedge, Northwest Territory sedge, slender sedge, and

River have the introduced Kentucky bluegrass, such as

water sedge. Sphagnum and philontus mosses occur

near the site of the old Buffalo Ranch in the Lamar Val-

with sedges in some areas. Wetlands with standing water

ley, along with native species that include Wheeler blue-

have hardstem bulrush and few-flowered spikerush,

grass, sheep sedge, and bearded wheatgrass.

with diamondleaf willow and Wolfs willow along the

Big sagebrush is found in the Yellowstone and Lamar

margin.24 In the riparian zone both meadows and shrub-

river valleys, where large numbers of elk and bison

lands occur. Riparian meadows typically have bluejoint

spend the winter; silver sagebrush and shrubby cinque-

reedgrass; tufted hairgrass; and other grasses, sedges, and

foil are found in meadows where the soils stay wetter

forbs. Shrublands along the creeks and rivers have alder,

for a longer time, such as in Hayden and Pelican val-

water birch, shrubby cinquefoil, silver sagebrush, and

leys (fig. 15.7). Associated plants include Idaho fescue,

various species of willow (diamondleaf willow, Geyers

junegrass, thickspike wheatgrass, and bearded wheat-

willow, sage willow, tealeaf willow, Wolfs willow, and

The Greater Yellowstone Ecosystem

Fig. 15.6. Mountain big sagebrush steppe and foothill grasslands are widespread in the valleys of the Lamar and Yellow
stone rivers, an area known as the northern winter range
of Yellowstone National Park. The forests and woodlands at

this elevation are dominated by Douglas-fir, many of which


have been killed by the Douglas-fir beetle. The grasslands are
dominated by bluebunch wheatgrass and Sandberg bluegrass.
Elevation 7,000 feet.

others).25 Dwarf willows (less than 6 inches tall) can be

thermal springs develop periodically, sometimes kill-

found above treeline, along with other alpine plants

ing trees over large areas.

common in the region (see chapter 14).


Several plant species are found only around hot
springs and geyser basins, including Ross bentgrassa

Grand Teton National Park

plant that is endemic to Yellowstone. The mineral-rich

Ferdinand Hayden returned to Yellowstone country in

water of these thermal features creates a saline envi-

1872 and sent a detachment, including photographer

ronment, and consequently, the characteristic plants

William H. Jackson, to explore Jackson Hole. The party

include such halophytes as alkali cordgrass, Baltic rush,

collected information on natural resources, took some

meadow barley, and seaside arrowgrass. The chemical

photographs, named many of the lakes and mountain

composition of the water varies from one geyser basin

peaks in Jackson Hole and the Tetons, and claimed to

to another, with the result that both terrestrial and

have been the first EuroAmericans to climb the Grand

aquatic species composition varies also. 26 For example,

Teton, though that has been disputed.27

the rock formed at Mammoth Hot Springs is rich in cal-

After 1872 the lands in YNP were off limits to set-

cium carbonate (travertine), but the rock around other

tlers, but Jackson Hole was still available. Named after

springs is rich in silica (sinter, also known as geyser-

early fur trader David Jackson, the town of Jackson

ite). Much of the color variation among geyser basins

was established in 1897.28 Cattle ranching became an

is caused by the varying temperature tolerances of dif-

important occupation in the valley at the same time as

ferent species of bacteria and algae, some of which are

potato farming developed in nearby Idaho at lower ele-

able to survive at near-boiling temperatures. New geo-

vations. To provide a more reliable source of irrigation

251

252 Landscapes of Special Interest

Fig. 15.7. Meadows and wetlands cover much of Hayden


Valley in Yellowstone National Park. The high point in the
distance is Mount Washburn, on the northern rim of the Yellowstone caldera. The valley is in the caldera and is underlain
by lake-bottom sediments deposited early in the Pinedale
glacial period. Moist meadows support mountain silver

sagebrush and shrubby cinquefoil; on drier sites, Idaho fescue


and Wyoming big sagebrush are common. Where the soil is
flooded during much of the summer, water sedge and beaked
sedge are common. The trees in the foreground are lodgepole
pine. Elevation 7,400 feet.

water, primarily for Idaho farmers, a dam on the Snake

on the mission of enlarging the park by buying as much

River was completed in 1906, thereby enlarging Jackson

land in Jackson Hole as possible and then donating it

Lake.29 Perennial concerns about possible dam failure,

to the federal government. 30 This was accomplished in

triggered by the 1976 failure of the Teton Dam 30 miles

1950, though not without criticism by many local land-

west (in Idaho), led to reinforcements on the Jackson

owners. As a compromise to enable park enlargement,

Lake dam (completed in 1989).

provisions were made for continued livestock grazing

The abundant wildlife and spectacular scenery south

and hunting. In 1972, the land along the highway con-

of YNP increased in popularity, and some began to

necting YNP and Grand Teton National Park was trans-

think that a second national park should be established.

ferred to the National Park Service by the U.S. Forest

Pierce Cunningham, a rancher near Spread Creek, cir-

Service and named the John D. Rockefeller Jr. Memorial

culated a petition in 1925 asking the state of Wyoming

Parkway, thereby acknowledging his vision.

or the federal government to set aside the valley for


the education and enjoyment of the Nation as a whole.
Four years later, Congress created Grand Teton National

Vegetation of Jackson Hole and the Tetons

Park. Predictably, the park boundaries encompassed

Jackson Hole is surrounded by mountains that extend

most of the rugged and uninhabitable parts of the

above treeline. Fifty-eight percent of Grand Teton

Tetons but almost none of the valley lying east of the

National Park is nonforested (alpine tundra, boulder

mountain frontyet it was the valley that needed pro-

fields, meadows, grasslands, and shrublands), 28 per-

tection from development. John D. Rockefeller, Jr., took

cent is lodgepole pine forest, 7 percent Engelmann

The Greater Yellowstone Ecosystem

tundra
Whitebark p.

Whitebark pine

Subalpine fir

Engelmann spruce

Lodgepole pine

Douglas-fir

Limber pine

Elevation
(feet)

Spruce-fir-pine

Lodgepole pine
Douglas-fir
Limber pine
sagebrush steppe

9000
8000

Fig. 15.8.Elevational distribution of


major forest types near Grand Teton
National Park. The solid vertical lines
show the range of elevations over which
each tree species is important as a forest
dominant; the vertical dotted lines indicate the range over which the species can
be found. Adapted from Baker (1986) and
Whitlock (1993).

7000
6000
5000

sprucesubalpine fir forest, 4 percent Douglas-fir wood-

common on the valley floor, especially on the glacial

land, and 1 percent aspen groves.31 In general, plant dis-

moraines where soil moisture is available late in sum-

tribution is determined by elevation, topography, and

mer (fig. 15.9). 33 Also, temperature inversions caused by

geologic substrate (fig. 15.8).

cold-air drainage may cause the valley bottoms to be

Douglas-fir and lodgepole pine are the most common

cool enough for lodgepole pine.

trees at low elevations in the Jackson Hole area, typically

Aspen groves are found in upland areas that are rel-

occurring together on Blacktail Butte and in the Gros

atively moist, such as at lower elevations where snow

Ventre Mountains to the east, where soils are deep and

accumulates or where there is groundwater seepage

fertile enough to support tree growth. Five Douglas-fir

to the surface. Frequently the groves are interspersed

habitat types have been identified, with the distinguish-

in sagebrush-dominated shrublands. Douglas-fir is

ing understory species being ninebark and Rocky Moun-

a common associate. 34 Mixed foothill shrublands or

tain maple on comparatively moist sites, and mountain

Douglas-fir grow on drier sites (see chapter 10). Sage-

snowberry, pine reedgrass, and spiraea on the drier sites.32

brush-dominated shrublands occupy the floor of Jack-

Lodgepole pine is common on some mountain slopes,

son Hole, except on glacial moraines, where conifer

such as Signal Mountain, as well as the west slopes of

forests are common (see fig. 15.9). Rivers and streams

the Tetons and the mountains to the north and east of

are fringed by riparian woodlands or shrublands (see

Jackson Hole, but it is infrequent on the east face of the

chapter 4). When overlooking the Snake River from

Tetons. At higher elevations on the Tetons, Douglas-fir

Antelope Flats, ancient river terraces are conspicuous

forests are replaced by subalpine forests dominated by

(see fig. 15.3), with coniferous woodlands on the steep

Engelmann spruce and subalpine fir. Whitebark pine is

slopes of cut banks and willow-dominated point bars

common above 8,000 feet, especially on dry windy sites,

on the opposite shore. Extensive willow shrublands also

and extends to upper treeline. Avalanche tracks on the

occur near Jackson Lake, such as the Willow Flats area

steep-walled canyons of the Tetons are obvious to hikers,

southwest of Jackson Lake Lodge.

extending from above treeline into the forested valleys

Much of the Teton range rises above treeline. The

and across streambeds, sometimes reaching a short dis-

steeper slopes appear barren, supporting only lichens

tance up on the opposite slope (see chapter 14).

and herbaceous plants in crevices. Alpine meadows and

In some areas Douglas-fir extends to higher eleva-

fellfields are found where soil has developed, such as

tions than does lodgepole pine, such as around Jackson

around Solitude and Amphitheater lakes. In parts of

Lake. Lodgepole usually occurs above Douglas-fir in

the Gros Ventre Mountains to the east of Jackson Hole,

the Rocky Mountains, but in Jackson Hole the pine is

sparse plant growth is seen at lower elevations, appar-

253

254 Landscapes of Special Interest

Fig. 15.9.This view of Jackson Hole from the top of Signal


Mountain shows forests of lodgepole pine and subalpine fir
on glacial moraines, and shrublands dominated by mountain

big sagebrush in the foreground and on the glacial outwash


plains. The Gros Ventre Mountains rise to the east beyond the
Snake River. See also fig. 2.6.

ently because of the shallow, infertile, often reddish

suggests relatively deep soils having considerable water-

soils that have developed there on Cretaceous shales.

holding capacity. The second type of community, found


farther south near the Jackson airport and Blacktail

Sagebrush Mosaic
Most of Jackson Hole is an unusually flat sagebrush
steppe, interrupted only by the Snake River valley and
a few buttes and moraines. Geologists refer to the area
as a glacial outwash plain, because it was deposited by
the meltwaters of the mountain glaciers that developed
80,000 to 12,000 years ago. One of the primary effects
of the meltwater was to wash away much of the silt and
clay that would have greatly enhanced the nutrient and
water storage capacity of the soil. Subsequently, however, this loss was compensated for by windblown loess
deposits from the west, such as from the Snake River
plains of present-day Idaho.
Close examination of the sagebrush steppe indicates
that at least three types of plant communities are represented. East of the Snake River, there is a dense cover

Butte, is characterized by an association of mountain


big sagebrush and bitterbrush. The abundant bitterbrush suggests a higher percentage of sand or gravel in
the soil. The third type of community can be seen west
of the Snake River in the southern part of the valley
and north of Moran. Here the sagebrush community
forms an intriguing mosaic, with patches of low sagebrush here and there where mountain big sagebrush
predominates (fig. 15.10). Low sagebrush is known to be
an indicator of shallow soils, sometimes with impeded
drainage or low fertility. Overall, the sagebrush mosaic
probably reflects different glacial outwash patterns as
well as differences in the rock types found in the mountains to the west and east of the valley. 35

Gros Ventre Landslide

of mountain big sagebrush, Idaho fescue, and other

The mountain ranges to the east of Jackson Hole have

species. The abundance of big sagebrush in this area

diverse originsthe volcanic Absarokas to the north-

The Greater Yellowstone Ecosystem

The forests and substrate slipped about 1,000 feet,


with many trees coming to rest at an angle. Only the
smaller trees survived, usually those less than 40 years
oldprobably because many of them were flexible
enough and light enough to avoid significant breakage.
Seedlings of subalpine fir, limber pine, aspen, cottonwood, and other trees have since become established
among the surviving Douglas-fir, Engelmann spruce,
lodgepole pine, and aspen. In contrast, the scar left by
the slide remains largely unvegetated.
Fears that the newly formed dam would break discouraged tourism. Indeed, 2 years later the upper 60
feet of the dam did fail, creating a flood of water, rock,
and mud that claimed the lives of six people. The

Fig. 15.11.The Gros Ventre landslide created Lower Slide Lake


when it dammed the Gros Ventre River on June 23, 1925. The
slide is more than a mile long and half a mile wide. The forests are dominated by lodgepole pine at lower elevations and
have lodgepole pine, aspen, Douglas-fir, Engelmann spruce,
and subalpine fir at higher elevations.

Fig. 15.10.East of the Snake River and north of Moran,


patches of mountain big sagebrush are found dispersed in a
shrubland of low sagebrush. Lodgepole pine and subalpine fir
dominate the adjacent forests on glacial moraines. Elevation
6,600 feet.

east, the gravel conglomerates of the Mount Leidy


Highlands, the sedimentary strata of the Gros Ventre
Mountains to the east, and the granite and gneiss of
Jackson Peak to the southeast. All are much older than
the Tetons. In the Gros Ventre Mountains, fine-textured
soils have developed that are highly susceptible to slippage when saturated. Landslides and mudflows are
common in this range. In fact, the largest slide in U.S.
history occurred there on June 23, 1925 (fig. 15.11).36
Despite a dense cover of coniferous forest at the time, a
block of earth a mile long, half a mile wide, and up to
300 feet deep slipped down the mountainside, crossed
the Gros Ventre River, and created a dam 900 feet long
and more than 200 feet high. Lower Slide Lake was created and now is about 3 miles long; ranch buildings
were inundated. All this occurred in about 2 minutes.

255

256 Landscapes of Special Interest

town of Kelly was swept away, and a 20-foot wall of

author was A. Starker Leopold, son of Aldo Leopold

water hit Wilson 2 hours later, 25 miles downstream.

the inherent dynamics of natural ecosystems were

A new alluvial fan was created in the Kelly area, but

explicitly recognized. 39 Earlier assumptions about the

the water level was back to normal the next day. Kelly

desirability of stable population sizes and vegetation

has been rebuilt at the same location, and new shrub-

condition were challenged. Instead of suppressing dis-

lands and riparian vegetation have developed over the

turbances, the parks were advised to accept and even

past 85 years. A trip to the Gros Ventre Slide provides a

encourage change driven by herbivory, predation, and

reminder of what could happen again, and is another

fire. This new natural regulation policy called for let-

illustration of how the geology of an area influences

ting such ecological processes operate to the greatest

plant distribution.

degree possible, with human intervention allowed as


needed for safety or to correct problems caused by pre-

Management Issues in the GYE

vious human activities. More research was also recommended, recognizing the many uncertainties about

Management concerns and priorities in the GYE have

whether specific ecological changes in the parks were

changed dramatically since YNP was established in

driven by natural phenomena or human interference.

1872.37 Initially, the objectives for YNP seemed simple

There also were debates about the long-term effects of

and straightforward: protect the geothermal features

Native Americans.40

from vandalism, protect the wildlife from predators

Whether natural regulation can really be an effective

and poaching, and protect the vegetation from fire

management approach in any ecosystem is still a topic

and overgrazing. Do these things and the park would

of considerable discussion. Consider, for example, the

be preserved. All fires, whether ignited by lightning or

challenges associated with the Yellowstone elk prob-

humans, were actively suppressed when possible; preda-

lem, the 1988 fires, the decline of whitebark pine, and

tors were shot; and elk and bison were managed inten-

the reintroduction of wolves.

sively, as though they were domestic livestock.


By the 1960s, however, it had become increasingly
clear that this approach to park preservation was overly

The Elk Problem

simplistic and ineffective. With much less predation,

By the late 1800s, unregulated hunting of elk and other

the elk population increased rapidly and appeared to

wild game drastically reduced herds throughout the

threaten the persistence of aspen, willows, and grass-

West.41 To counter this trend, animals were fed and

lands on their winter range. Fire-dependent species

predators were systematically eliminated, especially

were declining, and the iconic grizzly bear was on the

wolves. By 1912 the summer elk population in YNP

verge of local extinction. Throughout the parks his-

was crudely estimated at more than 30,000.42 Concern

tory some scientists had argued that the greatest value

shifted from fears of elk extirpation to worries that elk

of Yellowstone lay in preservation of its wild nature,

were damaging their winter range. In response, manag-

and had questioned programs of intensive control and

ers began trapping and shipping excess animals to other

manipulation of the parks ecosystem. For example,

places. When the demand for such animals declined,

Theodore Comstock celebrated the park as one of the

rangers reduced herd sizes by shooting the animals,

few remaining places where one could study the evo-

even inside the park.43 Still, elk numbers remained high,

lution and natural selection of animals in a state of

and concerns began to mount about overgrazing on the

nature, and George Wright and Aldo Leopold famously

northern winter range, located between Mammoth and

advocated protection and restoration of predators in

Cooke City. For example, ecologist Charles Kay com-

Yellowstone and other national parks. 38 Nevert heless,

pared historic and modern photographs, noting an

major changes in management philosophy were not

abundance of tall willows and a lack of browsing-caused

made until the 1960s, as proposed in a 1963 report

scars on aspen bark in the 1800s, but few willows and

titled Wildlife management in the national parks.

almost ubiquitous scars on aspen bark in modern views

Known as the Leopold reportbecause the senior

of the same locations (fig. 15.12).44

The Greater Yellowstone Ecosystem

every spring and fall between summer and winter


ranges. The summer ranges usually are at higher elevations, where plant productivity and forage availability
are greater because of the cooler, wetter environment.
Winter ranges are in the low-elevation foothills and
valleys (see chapter 10), where snow accumulation is
less and animals can forage until spring. However, the
relatively low productivity of winter range vegetation,
coupled with concentrations of animals for periods of
several months, leads to heavy use of forage plants
including shrubs and trees. Browse lines developed on
junipers, willows, and other tall shrubs (see fig. 10.16),
mature aspen were scarred after elk ate some of the bark,
and most aspen sprouts were snipped off before they
could grow into new trees (see figs. 10.14 and 10.15).
Recognizing that conversion of land to agriculture
and towns was reducing the availability of critical winter range for elk and deer throughout much of the West,
federal and state agencies initiated winter feeding programs to help maintain ungulate populations. In 1912
the National Elk Refuge in Jackson Hole was established
on the fertile bottomlands along Flat Creek. Wintering
elk began to congregate there by the thousands every
winter, as they do today.47 The Wyoming Game and
Fig. 15.12.The scars on these aspen were caused by elk feeding on the bark, a common occurrence on trees near winter
ranges.

Fish Department developed similar feeding programs


elsewhere. From the beginning, there has been controversy over the wisdom of this practice. The benefits are
perceived to be fewer agricultural conflicts, larger herds
for hunters, and less discomfort about animals starving

With the loss of tall willows came problems for

during harsh winters. However, the problems include

associated wildlife, including loss of nesting habitat

cost, injury of trees and other forage plants in the vicin-

for neotropical migrant songbirds.45 Competition with

ity of feeding grounds, disease transmission to livestock

elk for food was thought to be affecting populations

(such as brucellosis), and fostering the public perception

of deer and beaver. Some critics argued that few or no

that suitable habitat is not important.48

elk had occupied the winter range inside YNP until the

Arrival of the Leopold report in the 1960s led to

late-1800s, when land use changes north of the park

a fundamental rethinking of YNPs management

boundary restricted the animals to the park. To test this

approach. Some scientists argued that YNPs winter

hypothesis, historians Paul Schullery and Lee Whittle-

range was not overgrazed, pointing out that native

sey reviewed all observations made by explorers before

rangelands remained basically intact and continued

1882. They concluded that the record is not sufficiently

to sustain impressively large numbers of elk and other

detailed to know in any given year whether the elk

wildlife. Beginning in 1968, YNP terminated culling

population equaled, exceeded, or was less than at pres-

elk herds in the park, andas an experimentbegan

ent, but they became convinced that the park area was a

observing whether the Yellowstone elk population was

winter range for thousands of elk before 1882.46

capable of self-regulation. The park also implemented a

These concerns were not unique to YNP. In the GYE


and throughout much of the West, elk and deer migrate

research program to better understand the dynamics of


elk and their habitat.

257

258 Landscapes of Special Interest

Critics immediately challenged the natural regula-

mer ranges, which are more extensive than their winter

tion approach to managing the Yellowstone elk.49 They

ranges. These seasonal movements minimize grazing in

contended that YNP did not contain the key winter hab-

any particular area.54 Park biologist Douglas Houston

itat necessary to sustain a self-regulating elk population.

concluded in 1982, after 10 years of research, that reduc-

In winters since time immemorial, they argued, the

tions in the size of the elk herd were not necessary. 55

Yellowstone elk had migrated out of the relatively high-

In contrast to the grasslands, aspen and woody ripar-

elevation YNP to reach snow-free grasslands at lower

ian plants have continued to be significantly affected

elevations outside the current park boundary. Much of

by herbivores. 56 Approximately 85 percent of tree-sized

this area was now private land, converted to agriculture,

aspen on the northern winter range became established

roads, and other developments that impaired the qual-

prior to 1920; these old trees are gradually dying and

ity of the habitat for elk.

Hunting pressure in these

generally are not being replaced by young aspen sprouts

areas also reduced elk numbers significantly in severe

and seedlings.57 Willows and associated woody riparian

winters, and forced many of the elk to remain in the

vegetation in many areas appear degraded, with low

park during mild winters. Critics argued that regulation

stature, many dead branches, and new growth confined

of the elk population could never be truly natural, and

to the base of the plant rather than the top; seedlings

that unnatural overuse of the vegetation inside the park

of willow and cottonwood are often observed, but few

was unavoidable. To address this issue of elk migrating

survive.58 The National Academy of Sciences committee

out of the park during severe winters, wildlife agencies

evaluated several alternative hypotheses to explain the

and conservation groups eventually began to purchase,

lack of widespread aspen and willow regeneration dur-

or protect in other ways, some of the key elk wintering

ing the twentieth century. The committee determined

areas along YNPs boundaries.

that neither twentieth-century climate conditions nor

50

Notably, the assumption that wintering elk were

fire suppression were sufficient explanations in them-

irreparably damaging the vegetation of YNPs northern

selves for the paucity of recent aspen and willow regen-

range has not stood up to research findings. In 1998 the

eration and concluded that ungulate browsing was

U.S. Congress directed the National Academy of Sciences

the primary mechanism.59 Nevertheless, even though

to review all available science related to the manage

browsing is suppressing aspen and willows, the panel

ment of ungulates and the effects of ungulates on the

of scientists concluded that none of these species was in

vegetation of YNPs northern winter range, to clarify

imminent danger of extirpation on the winter range.60

what is known and what is not known. A committee of

Following cessation of artificial regulation in 1968, the

twelve eminent scholars concluded that the condition

northern Yellowstone elk population increased rapidly,

of the northern range is indeed different today than

growing from fewer than 4,000 animals in 1968 to more

when Europeans first arrived in the region, but that the

than 19,000 by 1988 (fig. 15.13).61 The controversy over

51

ecosystem is not in imminent danger of losing major

elk management in YNP and elsewhere did not subside,

components or crossing any threshold that would lead

but the plot thickened. In 1988 the largest fires in more

to irreversible changes. The committee determined that

than a century burned approximately a third of YNP,

grasslands on the northern winter range generally show

including a large portion of the northern winter range

less dramatic impacts of ungulate use than do aspen for-

and surrounding lands. With ongoing climate change,

ests and willow communities.52 Despite the consump-

the habitat has changed because of drought, a widespread

tion of 45 percent of the forage biomass by elk, the plant

epidemic of bark beetles, and the decline of whitebark

growth in grazed areas of YNPs northern winter range

pine due to an exotic disease. And then, in 1995, the gray

was 3685 percent higher than in exclosures, suggest-

wolf was reintroduceda major predator of elk. The so-

ing compensatory growth53 (see chapter 6).

called elk problem now became part of the broader chal-

One reason heavy grazing has had such minimal


adverse effects on the herbaceous vegetation of the win-

lenge of managing fire, forests, and wildlife as a whole.

ter range is that most foraging occurs during fall and

Fire Management

winter, when the plants are dormant. In spring and

Systematic efforts to eliminate wildfires were initiated

summer, elk and bison move to higher-elevation sum-

in YNP with the arrival of the U.S. Army in 1886,62

NUMBER OF ELK (thousands)

The Greater Yellowstone Ecosystem

25

Northern Yellowstone Elk Herd

20
15

1988
Fires

10
5
1970

1975

1980

1985

1995
Wolf
reintroduction

1990

YEAR

1995

2000

2005

2010

Fig. 15.13. Population trends in the northern Yellowstone


elk herd from 1971 to 2010. Beginning in 1968, the park
ended its program of artificially controlling elk numbers.
For that reasonand because of a series of moderate winters
and low levels of predationthe population grew. Large fires
burned approximately a third of the park in 1988, including
much of the northern winter range, and wolves were reintroduced to the park in 1995 after nearly a century of absence.

The recent decline in elk numbers is due to several factors, as


discussed in the text. Three sets of data are shown (black and
red points): Schullery and Whittlesey (1992) for 19711992
(black); Vucetich et al. (2005) for 19852004 (red), and the
Yellowstone Center for Resources (unpublished) for 2005
2010 (red). The data overlap in the period 19861992 and
differ slightly, reflecting the difficulty in censusing wildlife
populations.

but they were generally ineffective in remote forested

mented in surrounding wilderness areas and in Grand

areas. By the time a fire was reported and a crew found

Teton National Park. However, in 1988 researchers and

it, often days later, the fire commonly had either gone

managers learned that 15 years was not long enough to

out or had become so intense that it could not be extin-

fully understand the ecology of fire in the GYE.

guished with hand tools. Only with the availability of

The summer of 1988 turned out to be one of the

new fire-fighting equipment after World War II, notably

driest and windiest on record. Some 248 fires were

aircraft and smokejumpers, were managers able to sup-

ignited in the GYE by lightning or humans and burned

press most ignitions.

approximately 1.4 million acres (fig. 15.14). A massive

In the wake of the 1963 Leopold report, researchers

fire-fighting effort25,000 fire-fighters and $120 mil-

began reporting that fire was an important ecological

lionsucceeded in minimizing loss of life and property

process in most western forests and rangelands and

but could not stop the spread of fires across the land-

that fire exclusion was leading to undesirable ecologi-

scape until significant precipitation finally arrived in

cal changes. Therefore in 1972in the spirit of natural

mid-September.64

regulationpark managers initiated a cautious pro-

Park managers convened a panel of experts to eval-

gram of allowing lightning-caused fires to burn with-

uate the ecological impacts of the fires and to predict

out interference in a remote portion of the park during

likely ecosystem responses and management implica-

moderate weather conditions, when the risk of human

tions.65 The panel concluded that the size and severity

injury or serious resource damage was minimal. In 1976

of the fires were influenced more by drought and wind

this natural fire program was expanded to encompass

than by fuels or previous management practices. They

most of the parks backcountry area. During a period of

noted that large and severe fires also had occurred in

15 years, from 1972 through 1987, a total of 235 light-

the early 1700s (fig. 15.15).66 Subsequent research has

ning-ignited fires were allowed to burn, of which only

shown that large fires have recurred in Yellowstone at

27 burned more than an acre; the largest burned 7,400

intervals of decades or centuries throughout most of

acres.63 YNPs natural fire program seemed to be suc-

the Holocene.67 The 1988 fires were not confined to the

cessful, and similar natural fire programs were imple-

park but burned substantial portions of surrounding

259

260 Landscapes of Special Interest


Fig. 15.14.Extent and severity of the
1988 fires in Yellowstone National Park.
Burn severity was calculated using
data from pre- and post-fire Landsat
Thermal Mapper images and ground
observations. Especially striking is the
heterogeneity of burn severity. Similar
heterogeneity probably characterizes all
large forest fires. Cartography by Brian
J. Harvey. Reprinted with permission
from Romme et al. (2011, fig. 1) and
Springer Science and Business Media.

national forests and private lands, destroying some tim-

were too close and too flammable, they argued that

ber resources and threatening human life and property.

manager-ignited fires, during moderate weather con-

Even at 2.2 million acres, YNP was not large enough

ditions and especially near human habitations, could

to contain a natural fire regime. Moreover, the North

have minimized the anxiety of 1988. Analyses by For-

Fork Fire, one of the largest individual fires of 1988, was

est Service ecologist James Brown substantiated this

ignited by humans outside YNP and spread eastward

claim, suggesting that the cost of fuel reduction near

into the park, burning about 490,000 acres.68

developments would have been reasonable compared

Just as with elk management, the 1988 fires focused

to the cost of fighting the 1988 wildfires.69 However,

attention on the question of whether natural regula-

the same analysis led to the conclusion that the scale

tion can really be an effective management approach,

of the 1988 fires would not have been reduced signifi-

or whether a more active program is called for, even in

cantly had there been a program of prescribed burns in

the largest parks and wilderness areas. For some observ-

YNP. Human-caused prescribed fires would have been

ers, the need for manager-ignited prescribed fires was

set only under moderate fire-weather conditions, with

eminently clear. Noting that gateway communities and

the result that only small areas would have burned. The

visitor centers were threatened because adjacent forests

resources needed to contain such fires if weather condi-

The Greater Yellowstone Ecosystem

Complete
Suppression
Policy

Natural
Fire
Policy
LP3

40
% Area
in Each 30
Cover 20
Type
10

LP2
LP1
LP0

% Area 40
Burned 20
1700 20

40

60

80

1800

Fig. 15.15. Percentage area burned per decade from the


1690s through 1980s in a 450-square-mile landscape in central Yellowstone National Park, and percentage area covered
by four post-fire stages of forest development during that
time. LP0 is recently burned forest (typically 040 years
old), LP1 is young lodgepole pine forest (40150 years old),

20

40

60

80

1900

20

40

60

80

LP2 is mature lodgepole pine forest (150250 years old),


and LP3 is old-growth lodgepole pine forest, mixed with
spruce and fir on more fertile sites (250 or more years old).
Drafted by Michael Turner. Reprinted with permission of
M. G. Turner et al. (2001, fig. 7.15) and Springer Science and
Business Media.

tions and fire intensity had changed quickly would not

Natural fire policies in YNP and other national parks

have been sufficient to deal with more than a few fires

and national forests across the West were suspended

each year, and not every year would have had accept-

after the 1988 fires, while another review team evalu-

able weather conditions. Thus, assuming that human-

ated the scientific foundation of fire management.

caused prescribed fires had been used after 1972, when

The team concluded that YNPs natural fire policy was

fire was officially accepted as a necessary process in

basically sound but needed more specific criteria for

YNP, too little of the forest away from human habita-

deciding whether to suppress or to only monitor spe-

tions would have been burned to prevent the large-scale

cific fire ignitions. Yellowstone managers incorporated

1988 fires.

these changes in a new fire management plan in 1992,

Even if the 1988 fires cannot be viewed as entirely

and Grand Teton National Park and national forest

natural in size or spatial pattern, the ecosystem has

managers in the GYE developed similar modified poli-

shown great resilience, with surprisingly rapid natural

cies. Lightning-ignited fires are again allowed to burn

recovery of almost all components.70 There were some

in the GYE backcountry if they do not threaten life,

exciting ecological surprises as well: a large cohort of

property, or significant resourcesand if weather and

aspen seedlings appeared in burned areas the first year

fire-fighting resources are deemed such that another

after the fire (fig. 15.16). Genetic analyses confirmed

1988-like event is highly unlikely.

that the plants were genuine seedlings, not root sprouts,

But the Yellowstone fire story may have only just

Moreover, no aspen

begun. The 1988 Yellowstone fires were among the first

seedlings were observed in unburned areas, suggesting

of what has proven to be a dramatic increase in large

that successful sexual reproduction of this charismatic

forest fires across the American West since the mid-

Rocky Mountain tree may occur only in conjunction

1980s.73 The increase has been driven primarily by cli-

with severe fires. Another surprise was that, though the

mate changein particular, the trend of earlier spring

media sometimes gave the impression that YNP had

snowmelt, warmer summers, and a longer period each

been destroyed, visitation increased in the years after

year in which wildfires can burn. Anthony Westerling

the 1988 fires.72

and colleagues found a statistical correlation between

as are commonly seen in aspen.

71

261

262 Landscapes of Special Interest

fire rotation probably could not occur. Nevertheless, the


important message of this study is that fire will almost
certainly become more frequent in the future. With a
substantially shorter fire rotation, the older forests that
now characterize much of the GYE could not persist in
their present form. Moreover, post-fire vegetation development could be very different from the smooth recovery that occurred after 1988 if the climate is no longer
suitable for some of the major plant species.75 Westerling and his associates emphasized that there are many
uncertainties, but substantial changes in vegetation patterns appear likely.

Beetles, Blister Rust, Whitebark Pine, and Grizzlies


Whitebark pine is common at the highest elevations in
the GYE and Northern Rockies, where it stabilizes soils
and provides abundant seeds that are an important
food for grizzly bears and other wildlife. Recent fires
have burned whitebark pine forests, but this species is
capable of recovering from fire via seeds transported
and planted by Clarks nutcrackers.76 What the pine
may not recover from are the combined effects of global
climate change, mountain pine beetles, and a nonFig. 15.16.The small aspen in the foreground of this photo,
taken near Old Faithful, became established from seed after
the 1988 Yellowstone fires. These young aspen have been
heavily browsed by ungulates, but some have established new
clones in the area. The lodgepole pine in this photo are about
20 years old. The downed wood is from the trees killed by the
1988 fires.

native invasive pathogenthe white pine blister rust


(see chapter 14). Warming has allowed mountain pine
beetles to expand upward into whitebark pine forests
(see chapter 11 for beetle biology). Only recently have
entomologists learned that this pine lacks most of the
chemical defenses against bark beetles that are found
in lodgepole pine, most likely because the cold climate

large fires and warm summers in the Rockies from 1972

in whitebark pines high-elevation habitat formerly lim-

to 1999; they then applied this relationship to predict

ited bark beetle development and survival. When out-

fires for the next century, using four different global

breaks did occur in the past, they were infrequent and

climate change models.74 The results were startling: by

brief. However, the warm temperatures of the past 20

about 2050, large fires could be occurring almost every

years have allowed mountain pine beetles to success-

year, and the fire rotation in most of the GYE could be

fully attack whitebark pine over large areas; mortality of

reduced from the historical rotation of 120300 years

mature trees has been high.77

to less than 20 years. Fire rotation is a standard measure

The beetles generally have not killed the smaller

of fire frequency and is defined as the number of years

whitebark pine trees because, instinctively, the insects

required for cumulative area burned to equal the size

fly toward the larger trees that have adequate phloem

of the entire area of interest. This modeling procedure

to sustain a brood of beetle larvae. Ordinarily one

did not include changes in fuels; it assumed that fuels

would expect whitebark pine forests to recover from

would always be sufficient to permit burning even at

the beetle outbreak through growth of the smaller sur-

very short intervals. However, fires would begin to be

vivors. However, the white pine blister rust, a fungus

limited by fuel availability as intervals between fires

native to Eurasia that was accidentally introduced to

became ever shorter; the researchers noted that a 20-year

North America around 1900, kills whitebark pine trees

The Greater Yellowstone Ecosystem


Fig. 15.17.The grizzly bear, an omnivore once common across the Great
Plains and throughout the Rocky Mountains, is now found only in the Greater
Yellowstone Ecosystem and northward.
Photo by Mark Gocke.

of all sizes, large and small (see fig. 14.5).78 In some

ignated a threatened species in 1975. Whitebark pine

areas, blister rust infection appears to make the larger

seeds had been an important late summer and autumn

trees more vulnerable to mountain pine beetle attack,

food for grizzly bears, contributing to the accumulation

probably by reducing their ability to produce sufficient

of fat reserves necessary for winter survival. Because

amounts of resin.79 Thus, a warming climate not only

individual trees produce large amounts of seed only

leads to increased burning of high-elevation forests, but

every 35 years, it was important that the bears have

it also permits native bark beetles to move into these

large numbers of trees over large areas, so that some

previously unsuitable areas. Meanwhile, the exotic blis-

seed is available each year.

ter rust kills many of the small survivorsall leading

In addition to a declining number of whitebark pine

to a grim prognosis for whitebark pine in the northern

seeds, grizzly bears have also lost another food source,

Rockies.80 Efforts are under way to locate and propa-

the native cutthroat trout in Yellowstone Lake. The cut-

gate blister rustresistant individuals, with the intent

throat trout are threatened by two other non


-native

of restoring whitebark pine woodlands.81 However,

specieslake trout and a microscopic myxozoan that

given the spatial extent and rapidity of whitebark pine

causes whirling disease. Lake trout were illegally intro-

decline, many timberline forests of the GYE and other

duced to the lake, apparently in the mid-1980s, and

parts of the Northern Rockies may become what ecolo-

they have proven to be voracious predators of the native

gist Jesse Logan refers to as ghost forests.82

cutthroat. Whirling disease, first recorded in YNP in

Extensive loss of whitebark pine has had a cascad-

1998, leads to abnormal growth of a fishs bony struc-

ing impact on other ecosystem components, notably,

tures, causing an inability to feed.83 Cutthroat trout

grizzly bears (fig. 15.17). Now rare in the contiguous

populations have plummeted, and spawning runs up

United States and restricted to remote areas, the grizzly

tributary streamswhere the bears formerly harvested

was once common throughout the Rocky Mountains

great numbers of trout every summerhave dwindled.

and across the Great Plains. Its current range is prob-

Unlike the cutthroat, the lake trout do not spawn in

ably less than 1 percent of what it was in the 1800s.

rivers and, in deep water, they are inaccessible to bears,

With fewer than 1,000 animals in Montana, Wyoming,

ospreys, pelicans, and other native wildlife that had fed

Idaho, and Washington combined, the grizzly was des-

on the cutthroat.84 Park managers have implemented

263

264 Landscapes of Special Interest

an aggressive program of lake trout control, but at best,

matically from approximately 19,000 animals in 1994

the program manages merely to hold down the lake

to only about 6,000 in 2010 (see fig. 15.13). However,

trout population: eradication of this non-native invader

a recent assessment determined that wolf predation by

appears unlikely.85 Adding further to the bears troubles,

itself was not sufficient to account for the drop in elk

a third important food source for grizzlies has been

numbers. Contributing to the decline were increased

diminished since the reintroduction of wolves in 1995:

predation of elk calves by grizzly bears, as described

the bears formerly fed extensively on winter-killed elk

previously, along with drought and elk hunting outside

and bison when they emerged from their dens in the

the park.95 The wolves preyed mainly on elk, and occa-

spring, but now the wolves are consuming many of

sionally bison, deer, pronghorn, and moose, but hun-

those carcasses.86

dreds of cattle and sheep were taken as well (though

Surprisingly, despite the decline of three key food

less than 3 percent of livestock losses in the area can

resources, grizzly numbers in the GYE have increased

be attributed to predation). Compensation programs for

over the past three decades.87 The bears have proved

livestock losses were initiated by a nongovernmental

themselves to be genuine opportunists, shifting to

organization, Defenders of Wildlife, and then adopted

alternative resources when a preferred food source is

to varying degrees by state game and fish departments.

in short supply. The grizzlies are now killing more elk

The costs of conserving biological diversity can be sub-

calves than they did before the decline of cutthroat

stantial when large predators are involved, but the ben-

trout, thereby obtaining the protein and fat they need.88

efits are also obvious. For example, wolves and grizzlies

The bears also feed on berries, succulent plants, fleshy

are a major tourist attraction.96

roots, insects, and mushrooms when they are avail-

There also are thought to be ecological benefits. Fol-

able.89 Yellowstones grizzlies appear to be coping so far

lowing the sharp decline in elk numbers on Yellow-

with the effects of climate change and non-native spe-

stones northern range, ecologists William Ripple and

cies, although there surely are limits to what they can

Robert Beschta proposed that the associated reduction

endure.

in browsing pressure had allowed sprouts of aspen, willows, and other woody riparian plants to grow tall again
after nearly a century of ungulate-caused suppression.97

Wolves, Elk, and Aspen

Their suggested mechanism involved not just the reduc-

By the early twentieth century, the wolf was function-

tion in elk numbers but also a change in elk behavior.

ally extinct across most of the West, and it was one

With wolves, they postulated, the elk moved around

of the first species to be protected after passage of the

more often, avoiding habitats where they would be

Endangered Species Act in 1973. The wisdom of preda-

more vulnerable to wolf predation. According to Ripple

tor eradication in YNP had been questioned as early as

and Beschtas hypothesis, such movements might be

the 1930s, and the possible role of the wolf in control-

in response to detection of a wolf in the distance. In

ling excess numbers of elk was an issue in the 1960s.91

both cases, the result would be less time spent browsing

Wolf reintroduction gained considerable public support

in any particular place, allowing more sprouts to grow

at that time, but it was not until 1995 that 14 wolves

tall. Though an intriguing concept, other research has

were captured in Alberta and released into YNPs north-

not supported this behavioral hypothesis.98 Many aspen

ern winter range.92 In 9 years the population grew to 170

sprouts on the northern range still have not elongated

wolves in 15 packs distributed throughout the GYE. By

into new tree-sized stems, except within fenced exclo-

2012 there were several hundred in the area (about 100

sures where elk could not reach them.99

90

93

in YNP).

Unlike aspen, willows have become conspicuously

Critics worried that wolves would reduce elk herds to

taller since about 2000 in some well-watered places,

the point of impairing big-game hunting opportunities,

such as along the Lamar River near Soda Butte Creek,

and that they would prey heavily on livestock outside

perhaps due partially to the presence of wolves. How-

the park.94 The elk population did, in fact, decline. For

ever, increased willow growth has not been observed

example, the northern Yellowstone elk herd fell dra-

everywhere. Colorado State University scientists Kristen

The Greater Yellowstone Ecosystem

Marshall, Thomas Hobbs, and David Cooper concluded

underscored the importance of conservation efforts on

that a lack of willow and aspen food sources in the early

private lands in and surrounding the GYE. For example,

twentieth century led to local extirpation of beaver along

ecologists Andrew Hansen and Jay Rotella found that

some small streams, where beaver dams formerly main-

hotspots for bird diversity and abundance tend to occur

tained high water tables. Subsequent down-cutting in the

not in the high-elevation wilderness areas, but on low-

absence of beaver caused a lowering of the water table,

elevation private lands that are increasingly vulnerable

resulting in habitats that were too dry for willows, and

to degradation.102 In fact, this seems to be true for the

now, they reasoned, the willows in such places do not

region as a whole.

grow well even when fenced to exclude elk browsing.100

Given the natural complexity of this ecosystem

As another example of cascading ecological effects,

coupled with issues of boundary effects, land use

wolf reintroduction has led to a decline in the number

changes, invasive species, and climate changeit is

of coyotes in Jackson Hole, which in turn has resulted

difficult to know what management actions should be

in greater survival of pronghorn fawns.101 The value of

adopted, and to what degree natural processes are suf-

large wildlands for understanding ecological processes

ficient for maintaining the GYE.103 Managers are still

is again eminently clear.

learning about the effects of their decisions, and often


the lessons are unexpected. Historian Paul Schullery

On some days, the GYE is tranquil with elk, bison,

wrote, one of the most important contributions science

moose, and pronghorn at rest in expansive meadows.

has made to management dialogues in the past 30 years

The rivers are clean, and the thermal features are fasci-

is to elevate the admission of uncertainty as a credible

nating curiosities for resting hikers on a warm summer

management stance.104 Despite difficulties and con-

day. However, such tranquility belies the long winters

troversies, the Yellowstone country remains a national

that plants and animals must endure, the forest fires

treasure, indeed a global treasure. With the return of

that periodically burn across the landscape, and the

the gray wolf, the GYE again has all of the species and

molten rock that exists close to the earths surface. The

more or less the same ecological processes that existed

challenges of protecting such unique places, over such

when YNP was established in 1872. Equally important,

a large area, are formidable. Much research has focused

away from the roads, the GYE elicits a powerful sense of

on the GYEs public lands, but recent analyses have

wildness and beauty.

265

The Black Hills,


Bear Lodge Mountains,
and Devils Tower
Chapter 16

The mountains of western South Dakota and northeast-

hands, while the head-gear of the horses was deco-

ern Wyoming, known as the Black Hills and Bear Lodge

rated with wreaths of flowers fit to crown a queen of

Mountains, originated at approximately the same time

May. Deeming it a most fitting appellation, I named

as the mountains to the west.1 The highest point is Har-

this Floral Valley.

ney Peak, at 7,242 feet above sea level. Fifty million years
of erosion have stripped thousands of feet of sedimentary material, exposing the more erosion-resistant granitic core. Erosion occurred largely toward the east, the
direction in which the two major rivers flow, the Belle
Fourche and the Cheyenne (fig. 16.1). Devils Tower, an
unusual geologic feature, is commonly thought to have
been the neck of a volcano, though there is debate about
its origin (fig. 16.2).
Native Americans lived in the Black Hills when the
Vrendrye brothers, two French explorers, traveled
through the region in the mid-1700s. Most likely they
were the first Europeans to cross the northern Great
Plains and see the Rocky Mountains.2 More than a hundred years later, in 1874, Lt. Col. George A. Custer came
to the Hills from Fort Abraham Lincoln, near Bismarck,
North Dakota. Traveling with 1,200 troopers, 1,000
horses, and 110 wagons pulled by mules, plus 300 head
of cattle, his mission was to reconnoiter a route to the
Black Hills and explore their interior. For his journal
entry on July 25, Custer wrote:

Another member of his expedition observed, Everybody was making bouquets. . . . Some said they would
give a hundred dollars just to have their wives see the
floral richness for even one hour.3
The colorful flowers that Custer and his men enjoyed
most likely included anemone, blanketflower, bluebells,
blue-eyed grass, geranium, nodding onion, shooting star,
wild iris, and yellow owls clover.4 Custer also had his picture taken in the Black Hills with a dead grizzly bear.5
Congress designated much of the area as a forest reserve
in 1897, and today, most of the Black Hills and Bear
Lodge Mountains are in the Black Hills National Forest.
The Black Hills have been an important source of
gold over the years, with North Americas deepest and
largest underground gold mine located in Lead, South
Dakota. It closed in 2002 after 126 years of operation.
Many tourists toured the mine each summer. Visitors
continue to be attracted to the Hills by Devils Tower
National Monument, Jewel Cave National Monument,
Mount Rushmore National Memorial, Wind Cave
National Park, the Mammoth Site in Hot Springs, and

Every step of our march that day was amid flowers of

a summer climate that is cooler than that on the sur-

the most exquisite colors and perfume. So luxuriant

rounding plains. The region also is an important source

in growth were they that men plucked them without

of wood, water, and livestock forage.

dismounting from the saddle. . . . It was a strange

Botanically, one of the most interesting features

sight . . . the men with beautiful bouquets in their

of the area is the assemblage of plants found there

266

The Black Hills, Bear Lodge Mountains, and Devils Tower 267

Fig. 16.1. Geomorphic features of the Black Hills. The elevation ranges from 3,140 feet at Rapid City to 7,242 feet at the
top of Harney Peak, the highest point in South Dakota. On

the west side, the elevationa at Sundance and Newcastle are


4,750 and 4,334 feet, respectively. Adapted from Strahler
(1969).

(table 16.1).6 As might be expected, many Rocky

widespread about 10,000 years ago, when the climate

Mountain species occur in the area, such as elkweed,

was cooler and more humid. Today the Hills provide

grouse whortleberry, heartleaf arnica, narrowleaf cot-

an environment where many of these plants can still

tonwood, Oregon-grape, and ponderosa pine. Their

survive. Annual precipitation is considerably higher

presence is best explained by the generally accepted

than on the surrounding plains, ranging from 18 to 30

conclusion that Rocky Mountain plants were more

inches (fig. 16.3).

268 Landscapes of Special Interest

Fig. 16.2. Devils Tower, flanked by ponderosa pine savanna


and mixed-grass prairie. The grassland can be highly flammable. Common grasses include blue grama, western wheatgrass, needle-and-thread grass, and little bluestem. Fisher et al.
(1987) estimated that the historic fire-return interval in this
area was about 14 years. Devils Tower is one of a dozen igneous

intrusions in the Black Hills and Bear Lodge Mountains. The


others include Inyan Kara Mountain, Little Missouri Buttes,
and Warren Peaks in Wyoming, and Bear Butte, Black Butte,
Crow Peak, Custer Peak, and Terry Peak in South Dakota. Elevation at the top of Devils Tower is 5,117 feet. Photo by Hollis
Marriott.

More puzzling is the presence of plants that typi-

the somewhat cooler, wetter, largely forested environ-

cally are found in the deciduous forests of eastern

ment of the Hills, the biodiversity of the region as a

North America, such as American elm, bloodroot, box-

whole would be lower. Notably, the enabling legisla-

elder, bur oak, hackberry, and hophornbeam, or species

tion passed by Congress for establishing Wind Cave

from the boreal forests that stretch across Canada, such

National Park was to protect bison, elk, and pronghorn

as bunchberry dogwood, Canada scurvyberry, paper

as well as the cave itself. 8

birch, and white spruce. At least some of the eastern


species could have migrated westward along the moist
tributaries of the Cheyenne River, where they still occur

Geomorphic Regions

in favorable habitats.7 The northern species may have

The uplift and subsequent erosion of the Black Hills

migrated southward during the cooler periods associ-

and Bear Lodge Mountains have led to the formation

ated with the advance of continental glaciers. The Black

of five roughly concentric geomorphic regions that

Hills were not glaciated, but the boreal plants persisted

strongly influence landscape patterns (see fig. 16.1).9

a short distance south of the ice front where the climate

On the perimeter is the Hogback Rim, composed of

was favorable.

Mowry shale, Lakota sandstone, Fall River sandstone,

The Black Hills and Bear Lodge Mountains are refu-

Minnewasta limestone, and other erosion-resistant sed-

gia for many forest plants and animals that are sepa-

imentary strata that were tilted upward as the moun-

rated from their main populations by long distances

tains developed. The rim is sharply defined to the east

across grasslands. Conservation biologists place great

and south but is more like a plateau on the west. Pon-

value on such places, because they help maintain

derosa pine savannas predominate on rocky soils, with

genetic diversity. Without the habitat provided by

grasslands on fine-textured soils.

Table 16.1. Some plants in the Black Hills representative of other floristic regions
Region/Common name

Latin name

ROCKY MOUNTAIN SPECIES


Narrowleaf cottonwood

Populus angustifolia

Ponderosa pine

Pinus ponderosa

Creeping barberry (Oregon-grape)

Mahonia repens

Elkweed

Frasera speciosa

Heartleaf arnica

Arnica cordifolia

Grouse whortleberry

Vaccinium scoparium

Richardsons geranium

Geranium richardsonii

GREAT BASIN SPECIES


Alderleaf (true) mountain-mahogany

Cercocarpus montanus

Big sagebrush

Artemisia tridentata

Skunkbush sumac

Rhus trilobata

Blanketflower

Gaillardia aristata

Bottlebrush squirreltail

Elymus elymoides

Prairie junegrass

Koeleria macrantha

Threadleaf sedge

Carex filifolia

EASTERN DECIDUOUS FOREST SPECIES


American elm

Ulmus americana

Boxelder

Acer negundo

Bur oak

Quercus macrocarpa

Common hackberry

Celtis occidentalis

Green ash

Fraxinus pennsylvanica

Hophornbeam

Ostrya virginiana

Bloodroot

Sanguinaria canadensis

Downy yellow violet

Viola pubescens

Red columbine

Aquilegia canadensis

Virginia creeper

Parthenocissus quinquefolia

BOREAL FOREST SPECIES


Paper birch

Betula papyrifera

White spruce

Picea glauca

Bunchberry dogwood

Cornus canadensis

Canada mayflower

Maianthemum canadense

Single delight (woodnymph)

Moneses uniflora

Twinflower

Linnaea borealis ssp. americana

SOUTHERN GREAT PLAINS SPECIES


Bush morning-glory

Ipomoea leptophylla

Crested pricklypoppy

Argemone polyanthemos

Fendler (purple) threeawn

Aristida purpurea var. longiseta

Sand sagebrush

Artemisia filifolia

Velvetweed

Gaura mollis

Sources: Based on Buttrick (1914), McIntosh (1931), and Wright (1970).

270 Landscapes of Special Interest

104

resistant sedimentary strata. The vegetation is predominantly forests and woodlands of ponderosa pine, with

35

grasslands at lower elevations. The hills north of Newcastle are part of this region.11

Belle Fourche

Wyoming
South Dakota

55
50
40

Sundance

35

40

45

In the western part of the Black Hills, the sedimentary


strata form a comparatively flat area known as the Lime-

Spearfish

60

stone Plateau, with an elevation of 6,200 to 7,200 feet.


The underlying Madison Limestone (see chapter 2) is

Lead

45
55 50

permeable to water, which has enabled the formation of

40

numerous caverns, including Wind Cave and Jewel Cave.


Some creeks disappear into the bedrock before emerging

50

Rapid
City

Newcastle

again, such as Spearfish and Boulder creeks in the north.


44

marily ponderosa pine forests and savannas, commonly


45
50

45

Hot
Springs

10

10 mi

15

15 km

sandstones, forming the deepest and most spectacular


canyon in the Hills. The vegetation of the plateau is pri-

48

Custer

Spearfish Creek has cut through 400 feet of Paleozoic

intermingled with bur oak woodlands in the north and


groves of white spruce in canyons or on north slopes.
Also found are grasslands and foothill shrublands. A few
of the grasslands still have native species, but most are
now dominated by smooth brome, timothy, and other
introduced plants.12
The fifth geomorphic region is the igneous, granitic

Ardmore
Fig. 16.3. Mean annual precipitation patterns (shown in centimeters) in the Black Hills and Bear Lodge Mountains. Divide
centimeters by 2.54 to obtain inches. Adapted from Boldt and
Van Deusen (1974).

core of the Hills, which has been exposed through the


erosion of sedimentary rocks. Known as the Central
Area, it has many pinnacles and steep slopes (fig. 16.4).
The elevation is mostly 5,0006,000 feet.13 Again, the
vegetation is predominantly ponderosa pine forest, with
occasional groves of white spruce and aspen in wetter
habitats. Several lakes occur in the Hills, such as Sylvan

On the interior side of the Hogback Rim is the Red

Lake, but dams on creeks have created all of them.

Valleya name that stems from the red shales of the

The Bear Lodge Mountains, located to the northwest

Spearfish Formation that give a characteristic red color

of the Black Hills, occupy a smaller area and have less-

to the fine-textured soils.10 Also known as the race-

well-defined geomorphic regions. The highest point is

track because of its oval shape, the valley was formed

Warren Peak at 6,655 feet, about 600 feet lower than

by the erosion of softer shales, siltstones, and sand-

Harney Peak. Except for the absence of white spruce and

stones underlying the more resistant rocks of the Hog-

a few other species, the vegetation of the Bear Lodge

back Rim. Grasslands predominated in the valley when

Mountains is similar to that of the Black Hills.

European immigrants arrived, as they do today where


the land has not been cultivated. The towns of Sundance, Spearfish, Hot Springs, and the western part of

Vegetation Mosaic

Rapid City are found in the Red Valley, which ranges in

Nowhere in Wyoming and adjacent states are ponder-

elevation from about 3,000 to 3,600 feet.

osa pine forests so extensive (fig. 16.5; see fig. 1.5). They

Further to the interior, above the Red Valley, are

occur on nearly all soil types and exposures, occupy

the Minnelusa Foothills, composed of harder, erosion-

more than 90 percent of the Black Hills and Bear Lodge

The Black Hills, Bear Lodge Mountains, and Devils Tower 271

Fig. 16.4. (above) The granitic Central Area of the Black Hills
has widespread forests of ponderosa pine, with groves of
aspen and white spruce along ravines and in other relatively
moist habitats. The red ponderosa pine on the left was invaded by mountain pine beetles the previous summer. This
photo was taken in 2010. Many more trees have been killed
by the bark beetles since that time. Elevation 5,0006,000
feet. Photo by Carl Christensen.
Fig. 16.5. (left) Ponderosa pine forests are common in the
Black Hills, the Bear Lodge Mountains, the Laramie Mountains, and on east slopes of the Bighorn Mountains. Common
juniper is a conspicuous shrub at higher elevations. This
photo was taken in the Black Hills. Elevation 6,400 feet.

Mountains, and can be classified into at least seven


habitat types.14 Rocky Mountain juniper occurs with
ponderosa pine on drier locations, such as on the Hogback Rim; and two shrubssnowberry and bearberry
are widespread associates in the Central Area and on
the Limestone Plateau. Similar forests occur on escarpments in the Chadron National Forest in northwestern
Nebraska and in Custer National Forest in southeastern
Montana. Where annual precipitation is higher, such as
in the northern Hills, ponderosa pine and bur oak form
a distinctive conifer-hardwood association (fig. 16.6).

272 Landscapes of Special Interest

pole pine and limber pine, both very common in the


Rocky Mountains to the west, are found in only one
or two localities in the Black Hills. Engelmann spruce,
subalpine fir, and Douglas-fir are not found there at all.
The Black Hills region has four kinds of deciduous
woodland: riparian, woody draws, bur oak, and aspen.17
The first, found in ravines and along rivers at low elevations, is dominated by boxelder, green ash, hackberry,
bur oak, American elm, and plains cottonwood.18
Woody draws are found nearby and have thickets of
chokecherry, American plum, currant, Woods rose,
hawthorn, western snowberry, red osier dogwood, and
various willows (see fig. 10.12). Historical photographs
suggest that these woodlands were more dense before
livestock grazing and farming began.
The third type of deciduous woodland occurs on the
upland at lower elevations and is dominated by bur oak
(see fig. 16.6). It occurs most often in the northern and

Fig. 16.6. Bur oak is found at lower elevations in the Bear Lodge
Mountains and Black Hills, sometimes growing under an open
canopy of ponderosa pine. This photo was taken on the east
slope of the Bear Lodge Mountains. Elevation 4,250 feet.

Common plants associated with the pine and oaks are


aspen, chokecherry, common juniper, mountain ninebark, paper birch, and white spruce.
In all localities, fire suppression has led to an increase
in the density of ponderosa pine, with highly flammable forests now found in many areas.15 Most of the
forests are second growth, having been subjected to logging for many years. An outbreak of the mountain pine
beetle has killed many of the trees during the past 10
years, but dwarf mistletoe is rare or absent.16
White spruce groves are found in canyons and
ravines, on north-facing slopes, and in cooler, wetter habitats at higher elevations (fig. 16.7). In especially humid locations, such as on the north slopes of
ravines, the tree branches sometimes are draped with
a lichen known as Old Mans Beard (Usnea cavernosa
and U. sorediifera). Spruce occurs on both igneous and
limestone substrates, commonly in close proximity
with ponderosa pine, aspen, and paper birch. Lodge-

Fig. 16.7. White spruce is common in the Black Hills at high


elevations and along valley bottoms with cool-air drainage. It
is rare in the Bear Lodge Mountains.

The Black Hills, Bear Lodge Mountains, and Devils Tower 273

Fig. 16.8. Lt. Col. George A. Custer and his troops camped in
this meadow in 1874 (near the confluence of Silver Creek with
Castle Creek about 1 mile north of Deerfield Reservoir on U.S.
Forest Service road 110). The meadow appears the same today
as it was then based on a comparison of this photo with photos taken by expedition photographer William H. Illingworth.
However, the ponderosa pine forests in the background had
become more dense by the time this photo was taken in 2013.
A higher tree density could be attributed to natural forest
development after a fire that occurred a short time before
Custer arrived, or, alternatively, more trees could be explained

by fire suppression during the twentieth century. Notably,


mountain pine beetles are now thinning the forests, as shown
in fig. 16.9. Aspen groves can be seen on the edge of the
forest; Castle Creek flows through the foreground with various kinds of willows along its bank, including Bebb willow.
The meadow is now dominated by introduced species (such
as timothy, smooth brome, and Kentucky bluegrass), along
with green needlegrass, stiff goldenrod, and numerous other
natives. Illingworths photos of this meadow and others can
be viewed in Progulske (1974) and Grafe and Horsted (2002).
Elevation about 6,000 feet.

eastern parts of the Hills, mostly north of French Creek.

on the fringes of meadows (fig. 16.8). Associated species

Common associates include hophornbeam, ponderosa

include hazelnut, bracken fern, and wild sarsaparilla.

pine, green ash, American elm, and western snowberry.

Aspen can invade meadows and grasslands if browsing

On Mowry shale, bur oak forms a savanna with few

or grazing is not heavy. Aspen woodlands sometimes

understory plants. Sometimes the oak occurs as a shrub

give way to white spruce or ponderosa pine through

under ponderosa pine. Because of its ability to sprout,

succession. As with oak, the aspen and birch are able

oak can become the dominant tree if the pines are

to produce root sprouts and regain dominance more

killed by fire or harvesting. In fact, the re-establishment

quickly than ponderosa pine after a fire or other distur-

of pine may be difficult in some areas because of intense

bance (see chapter 11).

competition from oak and other species.19

The grasslands of the Red Valley and the fringes of

Aspen is the characteristic tree of the fourth type of

the Black Hills and Bear Lodge Mountains are typical

deciduous woodland.20 It occurs throughout the Black

mixed-grass prairie, dominated by blue grama, green

Hills on relatively moist sites. Occasionally aspen grows

needlegrass, little bluestem, needle-and-thread grass,

with paper birch or bur oak. The aspen woodland often

sideoats grama, western wheatgrass, and a variety of

forms a narrow band between the pine or spruce forest

sedges and forbs. 21 Where the soils are sandier, other

274 Landscapes of Special Interest

species are common, such as big bluestem, Indian rice-

wood are left in place for the benefit of wildlife and the

grass, and prairie sandreed. Westward, big sagebrush

maintenance of biological diversity. Research has shown

becomes more common. Streamside meadows in the

that at least 23 birds and 10 mammals in the Black Hills

same area have prairie cordgrass, tufted hairgrass, wild

depend on dead trees for food, nest sites, roosts, perches,

iris, and a variety of sedges. Inland saltgrass is com-

and dens.25 Sometimes the pines are thinned to provide

mon where the soils are saline and frequently moist.

more light, water, and nutrients for the growth of other

The interesting interaction of prairie dogs and bison on

desired species, such as bur oak or aspen. As tree density

the mixed-grass prairie of Wind Cave National Park has

is reduced, the growth and diversity of grasses, forbs,

been discussed in chapter 6.

and shrubs usually increase.26

At higher elevations, grasslands are found on the

Some believe there are too many roads in the Black

Limestone Plateau and in the Central Area near Deer-

Hills and that wood harvesting has taken precedence

field Reservoir. Those on the Limestone Plateau have

over other resource values, such as outdoor recreation,

been classified as Black Hills montane grasslands and

watershed protection, and biological diversity. Limit-

are identifiable by the presence of three grassesprairie

ing vehicular access is always controversial, but, though

dropseed, Richardsons needlegrass, and timber oat-

some roads have been closed, road density (miles of

grass. 22 The montane grasslands also have an abun-

road per square mile) in the Black Hills is still the high-

dance of colorful forbs, as Custer described in 1874.

est of any national forest in the region.

Several very large grasslands occur in the area, such as

The Black Hills also provide an abundance of live-

Gillette and Reynolds prairies. Only a few of the grass-

stock forage. 27 Excessive grazing still occurs in some

lands are still dominated by native plants, as most have

places (see chapter 6), especially in riparian meadows

been partially or completely converted to hay grasses,

and shrublands. Stream valleys, where much of the

mainly timothy and smooth brome. Those still domi-

private land is located, also are a preferred location

nated by native species are valued highly by conserva-

for summer homes. With livestock grazing and homes

tion biologistss.

concentrated in riparian zones, the habitat for big

Foothill shrublands are common in the region, typically on the Hogback Rim and lower parts of the Lime-

game, sensitive species, and other kinds of wildlife is


diminished.

stone Plateau. Skunkbush sumac and Rocky Mountain


23

juniper are common throughout the area, mountainmahogany occurs in the southwestern part of the Hills,
and snowbrush and russet buffaloberry are common in
the northern foothills.

Fire Suppression, Tree Density, and Streamflow


Perhaps the most subtle and far-reaching effect of land
management during the past century has been fire suppression, which is known to increase ponderosa pine
density. As discussed in chapter 11, surface fires char-

Land Management Issues


Timber Harvesting and Livestock

acterized many ponderosa pine forests and woodlands,


occurring historically every 1025 years, depending
on topographic position, elevation, and slope exposure. Surface fires kill most young trees, but usually not

Tree growth in the ponderosa pine forests of the Black

the older trees because of their thick bark. Such fires

Hills and Bear Lodge Mountains is the highest in the

also maintain a more open forest with relatively low

region because of the longer, warmer growing season

amounts of fuel. 28 Suppression led to the accumulation

and comparatively high summer rainfall.24 Ponderosa

of such large amounts of fuel that normally easy-to-

makes up 95 percent of the harvested wood. Consider-

control fires became stand-replacing wildfires.

able attention has been given to timber harvesting pre-

Although surface fires are thought to have been char-

scriptions that minimize adverse effects on other forest

acteristic of pre-European ponderosa pine forests, the

values. For example, clearcutting is generally avoided,

journals of early explorers suggest that crown fires also

with selective or shelterwood harvesting preferred, and

occurred. 29 For example, in 1880 H. Newton and W. P.

some standing dead trees (snags) and coarse downed

Jenney wrote:

The Black Hills, Bear Lodge Mountains, and Devils Tower 275

The Black Hills have been subjected in the past to

neous and have fewer large trees. Their results support

extensive forest fires, which have destroyed the

the work of William Baker and his associates.36

timber over considerable area. Around Custer Peak

The apparent effects of fire suppression are various

and along the limestone divide, in the central por-

and wide ranging. To illustrate, tree density has clearly

tion of the Hills, on the headwaters of the Box Elder

increased, with trees often invading adjacent grasslands

and Rapid Creeks, scarcely a living tree is to be seen

and causing a concomitant reduction in the amount of

for miles. . . . Some portions of the parks and val-

forage for livestock, deer, and elk (fig 16.9).37 After long

leys, now destitute of trees, show by the presence

periods with high tree density, many understory plants

of charred trees and decaying stumps that they

die. In contrast, and predictably, thinning by timber

were once covered by forest, but generally the pine

harvesting leads to an increase in understory plant

springs up again as soon as it is burnt off, though

growth. Notably, a large number of tree seedlings often

sometimes it is succeeded for a time by thickets of

become established if thinning or some other kind of

small aspens.30

disturbance is followed by relatively wet weather the

Similarly, R. I. Dodge observed in 1876 that Throughout the Hills the number of trees which bear the marks

following spring. Tree density increases quickly if there


is no surface fire to kill the young trees (fig. 16.10).38

of the thunderbolt is very remarkable. . . . The woods are


frequently set on fire. . . . There are many broad belts of
country covered with tall straight trunks of what was
only a short time before a splendid forest of trees. Early
photographs also suggest a more open forest with many
dead standing trees.31
Lightning strikes are still a major cause of grassland
and forest fires in the region.32 Some areas might have
experienced a comparatively long fire-free period in
the 1800s, which would have enabled fuel accumulation sufficient for the extensive crown fires observed
by early explorers. Douglas Shinneman and William
Baker concluded in 1997 that frequent surface fires
maintained open savannas on drier sites, such as in the
southern Hills, but that forests at higher elevations and
in the northern Hills had less frequent, more dramatic
stand-replacing crown fires over areas of 20,000 acres
or more. Thus, its likely that the Black Hills and Bear
Lodge Mountains as a whole have a mixed fire-intensity
regime, with both kinds of fires possible, depending on
climatic and fuel conditions at the time.33 Native Americans surely had an effect on the fire regime as well (see
chapter 11). 34
Fires have been suppressed for such a long time
and timber harvesting has removed so much of the treering evidencethat reconstructing the original forest
structure and fire history using dendrochronology has
been difficult. However, ecologist Peter Brown and his
associates found evidence for a diversity of forests, with
savannas interspersed with high-density forests.35 They
concluded that the forests of today are more homoge-

Fig. 16.9. Young ponderosa pine are invading this meadow,


located to the south of the Castle Creek meadow shown in
fig.16.8. Above the meadow, mountain pine beetles are killing
many of the ponderosa pines, opening up the forest after a
long period of increasing tree density.

276 Landscapes of Special Interest

toward fire continues, fostered for many years by onesided Smokey Bear ads and by fires sometimes spreading onto private land from national parks and national
forests. Yet the negative effects of fire suppression have
moderated those views, and prescribed fires are now
common on both private and public lands.41
However, fire management plans are not easily implemented. Some forests have become so dense that uncontrollable crown fires develop easily. Often the labor of
tree thinning and removal is preferred (fig.16.11). Prescribed burning in grasslands also poses significant
challenges. For example, fires are usually kept small so
that they can be extinguished easily if weather conditions change unexpectedly, but the new regrowth of
burned grasslands typically attracts bison and cattle. If
the burns are too small, excessive grazing of the new
plant growth results.
Another challenge, in the Black Hills and other parts
of the Rocky Mountain region (see chapter 12), is the
large number of homes and communities dispersed
through most of the Hills. Managers hesitate to ignite
prescribed fires where homes are nearby. Such fires
almost always burn within the agreed-to limits of the
managers, and almost always they can be put out at
Fig. 16.10. Ponderosa pine seedling density can be high after
a period of weather conditions that favor seedling establishment. If surface fires occur frequently enough, the young trees
are killed before a highly flammable forest develops.

will. However, there are still risks. Consequently, prescribed fires are used less often than some managers
would like. And when they are ignited, when weather
conditions are favorable, the slow-moving fires usually burn only a small portion of the area that would

Increased tree density and forest expansion are also

benefit from burning. The conundrum is that overly

thought to have caused a reduction in streamflow, as

cautious prescribed fire causes higher risks of wildfire

evapotranspiration is higher in forests and woodlands

in the future.42

than in grasslands.39 Watershed managers know that

Fuel loadings are clearly an important factor con-

streamflow increases after tree cutting or burning (see

tributing to flammability in low-elevation ponderosa

chapter 12). The journals of early explorers suggest that,

pine forests, but weather is important as well. Peter

in the 1800s, there were more beaver, greater water stor-

Brown examined fire scar data from sites distributed

age capacity in the stream banks, more plant growth

throughout the Black Hills and found that the years

in widespread riparian meadows, and possibly more

with the most fires were associated with droughts.43 He

streamflow (but fewer flash floods).

also found that new seedlings of ponderosa pine were

40

The value of periodic surface fires for maintaining

more likely to become established during wet periods,

grasslands, savannas, and more open forests is now

with the most extensive recruitment associated with an

widely recognized. Among the benefits are reduced

extended wet period from the late 1700s to early 1800s.

fuels, a lower probability of hard-to-control crown

From this he concluded that even-aged cohorts often

fires, and more streamflow and forage production. Also,

become established because of a sequence of favorable

native plants are maintained, thereby reducing the

climatic conditions, not catastrophic fires or the initia-

likelihood of invasive plants. Negative public opinion

tion of fire suppression.

The Black Hills, Bear Lodge Mountains, and Devils Tower 277
Fig. 16.11.This stand of ponderosa pine
has been thinned and is now thought
to be less susceptible to crown fires than
before. Because the small trees had little
market value, most of them were piled
and burned.

Considering the Black Hills region as a whole, the

story plants increased in abundance, such as snowberry,

largest fire in recorded history occurred in and around

yarrow, sideoats grama, needle-and-thread grass, Cana-

Jewel Cave National Monument in August 2000, follow

dian horseweed, and cudweed sagewort; others became

ing an extended dry period. Known as the Jasper Fire,

less abundant, such as common juniper, little bluestem,

it was of mixed severity and burned about 84,000 acres.

and ninebark. Unfortunately but predictably, several

Some of the forest had surface fires, other parts had

introduced species became more common as well, most

intensive crown fires, and still other areas did not burn

notably, bull thistle, Canada thistle, leafy spurge, and

at all. A patchy mosaic was created, with tree mortali-

Russian and spotted knapweed.

ties of about 20, 50, and 100 percent in areas of low,


moderate, and high burn severity, respectively.44 New
tree seedlings are becoming establishedthough slowly

Mountain Pine Beetles and Flammability

where burn severity was high, probably because most

Land management issues associated with insects, dis-

of the ponderosa seed supply was burned. Some under-

eases, and invasive plants are the same as those encoun-

278 Landscapes of Special Interest

tered elsewhere in the region (see previous chapters).45

In fact, the combination of beetles, fire, climate change,

An epidemic of the mountain pine beetle has been of

and other factors over a large area will promote the evo-

great concern during the past 20 years, causing much

lution and development of a more sustainable condi-

speculation about increased fire risk and what land

tion, whether managers take an active role or not.

managers might have done wrong to bring about such


conditions (see fig. 16.9). As noted, for a century or more
fires have been suppressed when possible and the trees

Invasive Plants, Climate Change, and Sensitive Species

have become more dense. The Forest Service and timber

Periodically the managers of federal and state lands are

industry collaborated in harvesting much of the wood

required to evaluate the ecosystems under their juris-

that was produced, though often not at the rate they

diction. In the Black Hills and Bear Lodge Mountains,

would have liked. The industry commonly claimed

that includes the Black Hills National Forest, Devils

the side benefit of improving forest health. Others felt

Tower National Monument, Jewel Cave National Monu-

that such harvesting caused habitat degradation over

ment, Wind Cave National Park, and Custer State Park.46

an excessively large area, or diminished the appeal of

Such assessments have been completed recently and, as

the area to tourists. Once again, land managers found

expected, all units are concerned about the challenge of

themselves in the middle of debates that often involved

dealing with the consequences of a long history of fire

governors and congressmen.

suppression and mountain pine beetle epidemics. Tim-

Presently, the furor seems to have diminished. More

ber harvesting has been accelerated where that option

people are recognizing that, on one hand, managing

is reasonable, and prescribed fires are used when there

ponderosa pine forests with attempts to suppress all

has been time to do the required preparation in places

fires will lead to other problems. Moreover, more people

where homes and other structures are either absent or

understand that the insects and diseases prevalent in

can be protected.

the Hills are nativepart of the natural biodiversity. To

Federal and state managers also are increasingly vigi-

manage the forest so that insect epidemics did not hap-

lant about the threats associated with introduced inva-

pen, or so that large fires did not occur, would create a

sive plants. Numerous species are involved, including

manicured forest that never existed before and many

buckthorn, burdock, Canada thistle, cheatgrass, com-

would not like. Even more roads would be required.

mon mullein, field bindweed, houndstongue, Kentucky

How to resolve such issues calls for creative collabora-

bluegrass, leafy spurge, spotted and Russian knapweed,

tion and acceptance of inevitable losses.

smooth brome, tumble mustard, white horehound, yel-

Current approaches for attempting to solve such

low toadflax, and yellow sweetclover. Curiously, the

problems include forest thinning and slash removal,

managers of Wind Cave National Park have observed a

which reduces the fuel load and increases the vigor of

rapid increase in white horehound, an introduced, inva-

surviving trees. The original structure of some pon-

sive plant in the mint family that is found primarily

derosa pine forests is thereby restored, at least to some

in association with the diggings of black-tailed prairie

degree. Prescribed fires are sometimes used when

dogs. Native species are adapted to such habitats as well,

weather conditions suggest that the risk of a wildfire

such as scarlet globemallow and fringed sagewort, but

is minimal and there are no mountain homes or busi-

they often are at a competitive disadvantage with the

nesses nearby, unusual circumstances in the area. Live-

introduced horehound. Mowing and pulling by hand

stock grazing is encouraged for reducing the abundance

are commonly used to keep weed populations in check,

of fine fuels, which may have an effect if that biomass is

and prescribed burning has been used successfully to

high after a year or two of above-average precipitation.

control some species.47 Herbicides in national parks

In some places the spread of aspen is promoted by cut-

are used carefully and with monitoring to minimize

ting the pines, which adds diversity to the landscape

impacts on other natural resources.

and could slow the movement of pine beetles. Accom-

The interaction between white horehound and black-

plishing such management activities over an area large

tailed prairie dogs is especially interesting because,

enough to be effective in a timely manner is difficult.

unlike the native plants, the horehound is not palatable

The Black Hills, Bear Lodge Mountains, and Devils Tower 279

susceptible to predation, simply because the uneaten


plants provide more cover for stalking predators. If this
happens, the colony could be reduced in size, whether
from predation or because some of the rodents move
to places where more of the vegetation is palatable and
easily clipped. Horehound is a robust plant; clipping it
over large areas by prairie dogs consumes a considerable
amount of their energy.
Wind Cave National Park is one of two locations in
the area with sufficiently large prairie dog colonies to
support black-footed ferrets; the other is about 50 miles
to the east, in Badlands National Park and the adjacent
Conata Basin in Buffalo Gap National Grassland. The
ferret population at Wind Cave has been fairly stable
in recent years, with about 50 animals in 2011.48 If the
prairie dog population drops because of insufficient food
or other factors (such as sylvatic plague or canine dis
temper), or because of increased predation by other carnivores, then the ferret population will decline as well.49
The number of prairie dogs is difficult to estimate,
but it is relatively easy to determine the aerial extent
of active colonies. Notably, the size of the colonies at
Wind Cave National Park has been observed to fluctuate
Fig. 16.12.The dark area in this prairie dog town, located in
Wind Cave National Park, was caused by an herbicide applied
the previous summer for the purpose of removing white horehound, an introduced invasive plant that prairie dogs do not
eat. This photo was taken June 14, 2013. Managers anticipate
that native species will become re-established where such
control measures are taken, though the success of this treatment remains to be seen. See text for discussion of how white
horehound is thought to affect the animals adversely. There
are numerous prairie dog colonies in the park that do not yet
have horehound, as illustrated in fig. 16.13.

with weather conditions, expanding during relatively


dry periods.50 Prairie dogs concentrating in smaller
areas during wet yearsso they can clip the taller vegetation sufficiently to reduce predationcould explain
this pattern. An alternative, untested hypothesis is that
the number of prairie dogs, rather than their extent,
declines during wet periods because of more effective
predation during years when the plant cover is higher
which facilitates stalking by predators (fig. 16.13).
Considering that climate warming could lead to more
frequent droughts, prairie dogs may expand into larger

(fig. 16.12). Thus, prairie dogsalong with bison, elk,

areas of short vegetationassuming sufficient food is

and pronghornfeed primarily on the native species,

available and predation pressure does not increase as

thereby favoring the spread of horehound. Unless the

well (see chapter 6).

prairie dogs develop a taste for horehound, the animals

Bison and pronghorn populations are prominent

will suffer as their native food supply is reduced. Indi-

grazers in Wind Cave National Park, spending much of

rectly, this could have detrimental effects on the endan-

their time in the Red Valley portion of the park, where

gered black-footed ferret that was reintroduced to Wind

mixed-grass prairie is common. Both species are often

Cave National Park in 2007, not to mention the other

seen grazing on prairie dog towns, most likely because

species that depend on prairie dogs and their burrows

they find more nutritious forage there (see chapter 6).51

(see chapter 6). Moreover, if the horehound is not eaten

Pronghorn and bison population sizes in the park cur-

or readily clipped by the prairie dogsor removed by

rently are thought to be appropriate for the available

park managersthe prairie dogs may become more

rangeland, but, with the threat of increased drought

280 Landscapes of Special Interest

Fig. 16.13. A mixed-grass prairie with black-tailed prairie


dogs in Wind Cave National Park. The grass and other plants
are rather tall, an indication that the prairie dog population
was not large enough to keep up with plant growth at the
time this photo was taken. Normally the rodents clip the
vegetation low, both for food and, presumably, to reduce the

amount of cover for predators. The pale green plant is fringed


sagewort, which, along with scarlet globemallow, is more
abundant around prairie dog mounds and in other disturbed
areas. Both are native species; there are no invasive plants
in this colony. Missing from grasslands such as this are the
carcasses of large mammals, as discussed in chapter 6.

stemming from climate warming, there is concern that

easily jumped by elk. At 7 feet, the new fences are taller.

forage and water for the animals may be insufficient.

There is still no hunting allowed in the park.

Management options are a reduction in herd size or


supplemental feeding and watering.

An entirely different concern was identified in the


condition report for Devils Tower National Monument,

Elk are another large herbivore in the park, though

namely, a decline in riparian woodlands because of the

they are seen less commonly and are not attracted to

dam upstream on the Belle Fourche River, forming Key-

prairie dog towns. In recent years their population has

hole Reservoir. Constructed in 1952, the dam reduced

exceeded the number that managers think the park can

the frequency and magnitude of flooding, which is

sustainat least under current climatic conditions

important for seedling establishment of plains cotton-

without adverse effects on their habitat. To facilitate

wood. The older trees are now dying faster than they

achieving the goal of 400 elk, several hundred animals

are being replaced (see chapter 4). Small trees have

were recently herded out of the park through gates in

been planted to replace them, but survival has been

what is now thought to be an elk-proof fence. Generally,

poor. Also, the dam transformed a clear, fast-moving

such fences should be avoided around national parks,

river to a warm, sluggish stream with high turbidity.

but they were constructed originally to prevent the

The sediment load is partially attributable to livestock

bison in Wind Cave National Park from interbreeding

grazing along the bank outside the monument. As

with cattle. The genetically pure bison population

often noted, management activities outside a natural

in the park is highly valued by conservation biologists.

area can have significant effects on the insideand

The original fences were effective for bison, but were

vice versa.

The Black Hills, Bear Lodge Mountains, and Devils Tower 281

Appropriately, condition assessments for the federal

Hills is marginal habitat for this species and that it also

and state lands now consider the implications of cli-

could become less common.53 The magnitude of such

mate change. Reports indicate that mean temperatures

changes can be debated, but the trends are the same.

in the region have risen by about 2F during the twen-

In general, future forests are likely to be much different

tieth century, and models predict an increase of 512F

than todays.

during the twenty-first century. As usual, there is less

As trees die from various inevitable disturbances, it

certainty about precipitation, but even if it increases,

would be ideal if landscapes could be restored to a condi-

evapotranspiration will increase as well because of the

tion that maintains the populations of sensitive species,

higher temperatures and longer growing season, most

such as aspen. Will that be possible with a warmer and

likely resulting in a drier climate. All managers in the

drier climate? Can the changes that are occurring be

region wonder what the effects of a warmer, drier cli-

guided so that the evolution of the native species can

mate will be.

continue, as they have for millennia? In fact, some sensi-

The most pertinent study on the effects of climate

tive species may not be able to survive; others may immi-

change was done in 2006 by U.S. Forest Service scien-

grate if the new environment is more favorable for them.

tist Gerald Rehfeldt and his associates.

The forests of the Black Hills and Bear Lodge Mountains

52

After careful

analysis of the climate required by ponderosa pine in

have changed dramatically before (see chapter 2).

western North America, they applied several climate


projections for the twenty-first century and concluded

Overall, the challenges of resource management are

that, most likely, ponderosa pine forests will be much

formidable in a landscape that has been modified by a

less dense and widespread in the Black Hills and Bear

century of fire suppression, habitat fragmentation, fenc-

Lodge Mountains by 2060 and that the species will no

ing, and the adverse effects of invasive species. And now

longer be reproducing by 2090 in some places where it

managers are faced with the uncertain effects of rapid cli-

now seems to thriveassuming unmitigated warming

mate change. The prospect of such changes is worrisome

continues. Mature trees can tolerate considerable tem-

to many, but two guiding principles are fundamental:

perature and drought stress, but as they die from old

conserve the diversity of native plants and animals and

age, beetle epidemics, fire, wind, or timber harvesting,

maintain the productive capacity of soils. One aspect of

the establishment of new seedlings will become less fre-

the Black Hills and Bear Lodge Mountains that will per-

quent than now because of the changing climate. Simi-

sist is the value of such places as refugia for mountain

larly, although aspen is the most widely distributed tree

plants and animals in an area otherwise dominated by

in North America, some think that much of the Black

the species of lowland grasslands and shrublands.

Chapter 17

The Laramie Basin

Westward from the Great Plains, beyond the Black

in previous chapters. Here the focus is on the Laramie

Hills and the Laramie Range, are a dozen or more inter-

Basin, one of the highest and most studied basins in the

mountain basins (see fig. 1.2). Long before they formed,

region (figs. 17.1 and 17.2).

the region was the site of inland seas in which thick


sequences of sedimentary strata accumulated (see chapter 2). As mountain ranges formed, the uplifted sedi-

Early Human History

mentary rocks were susceptible to erosion by water, ice,

The earliest evidence of humans in the Laramie Basin

and wind. Gradually, stones, sand, and silt produced

dates to about 13,000 years ago, based on the presence

by erosion were transported into adjacent basins,

of Clovis projectile points.2 By that time the climate

where they formed new strata. Hard as it is to imag-

was warming and the glacier that had reached the basin

ine, the basinsnow uplifted to a higher elevation

floor near Centennial had melted back to the slopes of

were nearly filled with material derived from adjacent

Medicine Bow Peak.3 Caribou, mammoths, mastodons,

mountain ranges. Some of the eroded mountains were

muskoxen, saber-tooth cats, and other exotic mammals

buried with these sediments. Geologists reached this

roamed across the tundra-like basin floor (see chapter 2).

conclusion because most basins are drained by a river

Cooling occurred again from about 1600 to 1850the

flowing through a canyon in the resistant rock of the

Little Ice Agebut not enough for glaciers to reach the

bordering uplift. Such a topographic feature develops

foothills as they had during the Pleistocene.4 Despite

where, at one time, the rivers course was established

climate changes, numerous Paleo-Indian artifacts sug-

over the buried mountain range. Gradually, the canyon

gest a frequent if not continuous occupation of the land

was cut by the river. As it deepened, more of the basin

in the vicinity of the Big and Little Laramie rivers.5

sediments could be washed away, further exposing the

The nomadic Blackfoot, Cheyenne, and Arapaho

mountain.1 Also indicative of a basins geologic history

tribes occupied the basin when Jacques La Ramie

are buttes and mesas, the tops of which are remnants of

arrived, at the age of about 35. Not much is known

ancient floodplains (see chapter 2).

about this French-Canadian trapper, but he is thought

Aside from these geologic similarities, the basins vary

to have been the first Caucasian in the area.6 Laura

in terms of elevation, climate, predominant vegetation,

Burdick wrote that La Ramie, in 1819, traveled up the

water availability, predominant land uses, exposed bed-

river on which he was camped, later to be named for

rock, mineral resources, water availability, and wildlife

him, against the advice of others who warned him

abundance. Some of this variability has been discussed

of the hostile Indians. He never returned and an

282

Fig. 17.1. Laramie Basin, with the Medicine Bow Mountains


on the west side and the Laramie Mountains on the east.
Centennial Valley, south of the town of Centennial, is defined
by the Medicine Bow Mountains on the left and Sheep Mountain on the right. The elevations of Laramie, Centennial, and
Woods Landing are 7,165, 8,074, and 8,095 feet, respectively.
The elevation of Wheatland Reservoir at the north end of
the basin is 6,959 feet, 1,136 feet lower than the Laramie

River at Woods Landinga straight-line distance of about


60 miles. The highest point in the Snowy Range, that part of
the Medicine Bow Mountains from which glaciers originated
during the Pleistocene, is Medicine Bow Peak12,013 feet.
The red dashed lines show the location of two transects used
for the profiles of geological surface types and vegetation
types shown in figs. 17.15 and 17.16. Cartography by Ramesh
Sivanpillai, University of Wyoming.

284 Landscapes of Special Interest

Medicine Bow
Mountains

Feet above
sea level

Snowy Range

12,000

Laramie
Range

Laramie
Basin

7,000
Mesozoic
Paleozoic

West

30 miles

East

Scale

Qa

Laramie River

Qa

Feet above
sea level

Laramie Range

Laramie Basin

8,600

Qa
7,200

Qa

Ku

Pf
Ps

Jm

West

4 miles

East

Scale

EXPLANATION
Quaternary
Alluvial deposits
Mesozoic
Cretaceous formations, undivided
Morrison Formation (Jurassic)
Chugwater Formation and Jelm Formation (Triassic)
Paleozoic
Forelle Limestone (Permian)
Satanka Shale (Permian)
Casper Formation (Pennsylvanian-Permian)
Precambrian
Granite
Undivided metamorphic rocks
Fault, arrow indicates relative direction of movement
Shear zone

Fig. 17.2. Geologic structure of the Laramie Basin. The top diagram is a generalized cross section of the entire basin, showing
the underlying sedimentary strata and the fault that was important in the formation of the Medicine Bow Mountains. The
bottom diagram shows more detail in the vicinity of Laramie.
The abundance of Quaternary alluvial deposits on the surface
is indicated, as is the thickness of the Casper Formationan
important aquifer in the basin. The Forelle Limestone forms a
prominent ridge east of Laramie; the colorful Casper Formation is exposed south of Laramie in the vicinity of Red Buttes
and Chimney Rock. Figure 2.3 illustrates the geologic development of the basin; a geologic time chart is presented in fig. 2.1.
Adapted from Mears et al. (1986).

unfinished cabin was found near Sybille.7 It would be

herds of buffalo, antelope [pronghorn] and mountain

good if more could be learned about his life, consider-

sheep. He also wrote about the many small streams that

ing that so many places bear his nameLaramie Peak,

were bordered with a thin growth of small willows and

Laramie River, Laramie City, Fort Laramie along the

richly stocked with beaver. Narrowleaf cottonwood

North Platte River, and Laramie County to the east of

also was observed. Another explorer, E.Willard Smith,

Albany County.

entered the valley from the south in 1840, observing

In 1825, about 5 years after La Ramie arrived, the fur

a great many buffalo every day as we passed along.

baron General William Ashley led an expedition across

In 1843 Captain John C. Fremonts party was in the

the basin. The general wrote in his journal that the val-

basin with Kit Carson, killing a pronghorn near Tie

ley and mountains were enlivened by in


numerable

Siding and camping along the Laramie River. As they

The Laramie Basin

passed around the north end of the Medicine Bow

established, and just north of the Laramie Basin, near

Mountains, sagebrush became more common, and the

Rock River, small deposits of coal were mined. With

party continued to encounter bison. Francis Parkman,

the exception of limestone for Portland cement, Albany

author of The Oregon Trail, was in the basin in 1846

County and the Laramie Basin generally are considered

and observed wolves, beaver, antelope, sage-grouse,

mineral poor. The rivers, lakes, and nearby mountains

elk, bighorn sheep, mule deer, and a forest fire. Three

provide many opportunities for outdoor activities. In

years later, in 1849, Captain Howard Stansbury and Jim

2012, Albany County had 37,000 residents83 percent

Bridger crossed the basin with a party of 18 mounted

living in Laramie.

horsemen, experiencing altercations with indigenous


tribe members along the way. Grasslands covered most
of the basin floor, as they do now. Stansbury also wrote
about bison, and he commented on an isolated grove of

Glacial Outwash, Hogbacks, Red Buttes, Big Hollow,


and Other Geologic Features

cottonwoods along the Laramie River, near where the

During the Pleistocene, glaciers in the Medicine Bow

city of Laramie would be founded 19 years later.8

Mountains of Wyoming and Colorado loosened large

The pace of EuroAmerican immigration accelerated

volumes of rocks, stones, sand, and silt.12 When the gla-

with the establishment of the Overland Stage Line by

ciers melted, much of this material was flushed onto the

Ben Holladay in 1862.9 This trail entered the basin from

basin floor, creating deltas. The waterways transporting

the southeast, near present-day Tie Siding, and contin-

and depositing this sedimentthe ancestors of the Big

ued westward around the north end of the Medicine

Laramie River, the Little Laramie River, and their tribu-

Bow Mountains (often referred to as the Snowy Range).

tariesshifted back and forth across the basin. These

Travelers with carts and wagons followed this trail.

waterborne sedimentslabeled Quater


nary alluvium,

Indian attacks led to the establishment of Fort Sanders

because they have been deposited since the beginning

in 1866, the location of present-day Laramie. The cav-

of the Pleistocene, 2.6 million years agoare the parent

alry provided security for cattle ranchers and for the

material for soils over about 75 percent of the Laramie

construction of the first transcontinental railroadthe

Basin floor (fig. 17.3). As might be expected, more large

Union Pacific. The railroad arrived in 1868, the same

rocks occur near the foothills than farther to the east

year that the city of Laramie was founded. After a brief

and north, where the flow of water was less forceful.

period of lawlessness and citizen vigilantes, Laramie

Before and following glacial advances, there were long

was settled with churches, schools, banks, a hospital,

periods of erosion from the basinbut only as rapidly

a courthouse, and, in 1873, the Territorial Federal Peni-

as the Laramie River could cut a canyon through the

tentiary. With the proximity to the railroad and the

Laramie Mountains to the northeast, a process that con-

availability of water, land privatization was more com-

tinues to this day.

plete than in basins to the west, where much of the land

Hogback ridges are conspicuous topographic features

to this day is in the federal domain. In 1886, before

on the perimeter of intermountain basins where sloping,

statehood in 1890, the Territorial Legislature designated

layered sedimentary strata are exposed. Strata of rela-

Laramie as the site for Wyomings first and, still today,

tively soft shale and siltstone erode more rapidly than

only 4-year university.

those of harder sandstone and limestone, with the result

During the 1900s there was great optimism about

that the more resistant strata became ridges (see fig. 17.2).

the potential for various land uses in the Laramie

The ridges create a different environment for plant and

Basin.10 Irrigated croplands were envisioned, but the

animal life. Mountain-mahogany and occasionally lim-

growing season proved too short and cool; only about

ber pine and ponderosa pine are found on the rocky soils.

3 percent of the basin is cultivated, mostly for hay.11

The ridges also add aesthetic appeal to the landscape and

The raising of cattle and sheep has been the most viable

provide desirable habitats for wintering animals, espe-

form of agriculture. Limestone and gypsum deposits

cially on leeward or south-facing slopes that are warmer

supported a brick and plaster industry for a time, and

in the winter (see chapter 10). People also enjoy having

later a cement industry. Several small oil fields were

homes in such places.

285

286 Landscapes of Special Interest

Fig. 17.3. Surface geology for the Wyoming portion of the


Laramie Basin. Recent alluvium and alluvial fans, shown in
yellow and buff colors, cover about 50 percent of the basin
floor. Green and blue colors show terraces and benches, which
originated as alluvium and occur on about 24 percent of the
floor. The Airport and Harmony benches are conspicuous (see

also fig. 17.5). Much of the surface has remnants of the Quaternary alluvium and older deposits that once filled the basin.
Some surface types listed in the legend are not easily distinguished at the scale of this map, including eolian deposits.
Cartography by Ramesh Sivanpillai, University of Wyoming.
Adapted from Case et al. (1998).

Prominent on the east side of the Laramie Basin are

aquifer is recharged by rain and snow on the west slope

the hard rocks of the Casper Formation, strata deposited

of the range, from its upslope edge near Lincoln Monu-

during the Paleozoic Era and now easily visible in the

ment down to the Laramie city limits. Protecting water

canyon east of Laramie, through which Interstate 80

quality in this aquifer is viewed as very important.14

passes. It is the oldest sedimentary formation in much of

The erosion of Quaternary alluvium and underlying

the basin, lying directly on top of Precambrian igneous

strata has been most rapid in the southern half of the

and metamorphic rocks (see fig. 17.2). It includes strata

Laramie Basin, where the elevation is about 1,000 feet

that contain the aquifer on which most people in the

higher than on the north end. This erosion has gone on

basin depend. Wells can be drilled to access this water,

sufficiently long that some of the basins oldest sedimen-

but artesian flows reach the surface in numerous places

tary strata are now exposed, often creating landforms

as springs, such as on the east side of Laramie.13 The

that are aesthetically, economically, and ecologically

The Laramie Basin

basin within a basin, about 10 miles long and 2 miles


wide, is easily seen along Wyoming Highway 130 (see
fig. 17.5) and is thought to have been formed by wind
erosion, starting most likely when glaciers extended
into the foothills above the town of Centennial. Strong
winds, possibly of hurricane force, apparently developed because of steep temperature gradients from the
top of the ice mass down to the valley.15 Such winds
dried and scoured the soil over a long enough period to
break down the deposits of calcium carbonate that previously had cemented the surface. Fine materials might
Fig. 17.4.Cross-bedding in sandstones of the Casper Formation indicate that the sand in this area was deposited in a
coastal dune field that existed here about 300 million years
ago. This photo was taken near Chimney Rock in the higher,
southwestern part of the Laramie Basin.

have blown eastward over the Laramie Range, whereas


coarser particles would have been carried downstream.
The bottom of Big Hollow is 300 feet below the surface
of the Airport Bench and has a playa lake in the bottom.
The smaller Bamforth Basin to the north most likely was
formed in a similar manner and also holds a playa lake.
Except for just above the town of Centennial, glaciers

important. For example, colorful sandstones of the Cas

never extended onto the floor of the Laramie Basin. The

per Formation are easily seen south of Laramie in the

melting of the latest glacier in this area began roughly

vicinity of Red Buttes. The brick-red pillars and monu-

10,000 years ago and led to the formation of a large,

ments add aesthetic appeal to the terrain. An equally

lobe-shaped glacial outwash plain, around which the

picturesque feature, known as Chimney Rock, is a rem-

Little Laramie River flows, first southward and then east

nant of Casper Formation sandstone and the underlying

and northeast. This relatively recent alluvium is easily

Fountain Formation, about 25 miles southwest of Lara-

seen along Highway 130 from Centennial eastward for

mie. Distinctive for its mass and shape, this natural mon-

about 6 miles, to the second crossing of the Little Lara-

ument is more than 200 feet high and has been sculpted

mie River (see fig. 17.3). To the west of Centennial, the

by erosion along Sand Creek. Stabilized sand dunes exist

highway climbs onto a terminal moraine marking the

in the area, formed from sand first deposited hundreds

farthest advance of the glacier.

of millions of years ago. The distinctive cross-bedding in

Various minerals have been extracted from the rock

the Casper sandstone exposed in the vicinity of Chim-

formations underlying the Laramie Basin. For a time in

ney Rock is evidence that it formed in a windblown

the nineteenth century, gold and copper mining took

coastal dune environmentnear the edge of an inland

place around Centennial and a few other localities.

sea that existed about 300 million years ago (fig. 17.4 ).

The most important minerals in the past 50 years have

Other topographic features also illustrate the geo-

been gypsum from the Pennsylvanian strata exposed

logic history of the basin. For example, west of Laramie

at the south end of the basin and limestone from the

are two large, flat, elongated benchesHarmony Bench

Casper Formation, along the west flank of the Laramie

and Airport Bench (fig. 17.5). Like buttes and mesas in

Range. Much of the stone used in the construction of

the area, these benches are capped by remnants of for-

older buildings on the University of Wyoming campus

mer floodplains. Now, on both sides, they are under

is sandstone from the Casper Formation.

going slow erosion by wind and water (see chapter 2).

Fossil fuel extraction has been a minor part of the

The tops of the benches have an abundance of river-

Laramie Basins economy. Only small amounts of oil

bed cobble that armors the surface, slowing the rate of

have been pumped from several small well fields, and,

erosion.

as noted, coalbeds are essentially nonexistent. Most

In addition, there is Big Hollow, reputedly the sec-

of Wyomings coal comes from Paleogene (early Ter-

ond largest deflation basin in the world. This elongated

tiary) strata, formed when subtropical swamps existed

287

288 Landscapes of Special Interest

Fig. 17.5.Topographic features in the Laramie Basin west of


Laramie, with Sheep Mountain in the background. Highways
and the Big and Little Laramie rivers are shown. Note the
location of the Airport and Harmony benches, and the large
deflation basin known as Big Hollow, west of the Airport
Bench. Bamforth Lake is in a smaller deflation basin. Lake

Hattie was enlarged to store more water for the irrigation of


hayland in the Pahlow strath, a term used for the valley between the south rim of Big Hollow and the Harmony Bench.
Drawing by Brainerd Mears, Jr., courtesy of the Wyoming
Geological Survey, Laramie.

over much of the area (see chapter 2). Such rocks have

annual temperature from 1991 to 2010 at the Laramie

been eroded from the Laramie Basin because of its high

Regional Airport was 41F, with July average highs and

elevation.

lows of 80F and 49F, respectively (fig. 17.6).16 Annually, most of the basin had only 75100 frost-free days

Elevation, Climate, and a Double Rainshadow

during this time. The short growing season combined


with relatively cool summer temperatures limits tillage

With an average elevation of about 7,200 feet, the

agriculture to mostly hay production where irrigation

Laramie Basin has the coolest climate of all the basins

water is available, such as near the Big and Little Lara-

in Wyoming, though similar to the Washakie and

mie rivers and along Rock Creek just to the north of

Green River basins to the west (see fig. 3.7). The mean

the basin.

The Laramie Basin


4.0
3.2

66

2.4
43
1.6

32
19

-5

0.8

that moisture falls in the basin. The annual precipitaMonthly Precipitation (inches)

Monthly Temperature (F)

90

0.0

Fig. 17.6. Average monthly precipitation (bars) and average


high and low temperatures (curves) at the airport in Laramie
(19812010). Comparable graphs from other weather stations
can be seen at http://www.usclimatedata.com.

tion at Laramie is only about 11 inches, most of which


falls in April and May. The mountains and basins to the
west tend to receive a higher proportion of the annual
precipitation in the winter, usually more than half. As a
result, sagebrush steppe is more common in the basins
to the west (see chapter 7).
Even with low precipitation, the Laramie Basin
was hospitable to both indigenous tribes and Euro
Americans because of snowfall in the adjacent mountains, which fed the Big and Little Laramie rivers in
addition to numerous springs on the east side of the
basin. Low-lying areas had numerous lakes and ponds,
such as in the vicinity of Lake Hattie and in the Sand
Creek drainage (see fig. 17.5).17 Waterfowl were abundant. Notably, creeks flowing into the basin from the
drier Laramie Range are ephemeral and provide little

The Laramie Basin is also windy at times, with a

surface water.

mean annual velocity estimated at about 12 miles/


hour. For comparison, mean annual wind velocities
for Missoula and Denver are about 6 and 9 miles/hour,

Vegetation Patterns

respectively. Snow drifting commonly occurs in lee-

Forests and woodlands are conspicuous as one crosses

ward locations (see also chapter 3), which favors the

mountain passes and descends into the Laramie Basin

development of shrublands dominated by big sage-

from the east, south, and west, whereas shrublands are

brush. Grasslands are found on surfaces that usually

more likely to be seen when entering at a lower eleva-

are blown free of snow.

tion from the north. However, mixed-grass prairie cov-

The basin is semi-arid for two primary reasons.

ers about two-thirds of the basin floor (fig. 17.7). Plant

First, evapotranspiration is more rapid at high eleva-

cover is sparse on windswept knolls, consisting of little

tions, especially when the wind is blowing. Second,

more than a few cushion plants and arid-land sedges

annual precipitation is quite low because of mountain

(threadleaf sedge and needleleaf sedge). In more favor-

ranges to the east and west that cause rainshadow on

able environments, the vegetation is dense, dominated

both sides. Several times each season it is common

by blue grama, western wheatgrass, Sandberg bluegrass,

to see storm clouds on the horizons that never reach

junegrass, needle-and-thread grass, fringed sagewort,

the basin. Those on the east are typically produced by

sand lily, Hoods phlox, and many other species (figs.

upslope storms, created when a low-pressure cell passes

17.817.10). Big sagebrush, mostly Wyoming big sage-

through southern Colorado or northern New Mexico.

brush, usually grows where snow accumulates on the

At such times, moist air from the Gulf of Mexico cir-

leeward side of ridges. Less than 10 percent of the basin

culates counterclockwise onto the east slope of the

has sagebrush-dominated shrublands, and much of that

Rocky Mountains. As the air rises, the air cools and

is black sagebrush on shallow soils in the foothills.18

condensation occursthe upslope conditions men-

Mountain-mahogany grows on exposed bedrock, and

tioned by weather forecasters. The precipitation that

some mountain big sagebrush is found where the soil

falls is mostly on the east side of the mountains. To the

is deeper and more snow accumulates (see chapter 7).

west of the basin, the air rising over the Medicine Bow

In general, plant growth and species composition vary

Mountains generally flows from the Pacific Ocean and,

largely as a function of wind exposure; the potential

having already passed over several mountain ranges,

for snow drifting; and the depth, salinity, and water-

is comparatively dry. Thus, only a small proportion of

holding capacity of the soil.

289

Fig. 17.7. Land cover in the Wyoming portion of the Laramie


Basin. Grasslands occupy about 66 percent of the basin floor,
with shrublands dominated by big sagebrush and greasewood
occurring on about 10 and 5 percent of the land, respectively. About 9 percent of the basin is in riparian woodlands,
shrublands, and meadows; about 3 percent of the land is in
cropland agriculture, half of which is irrigated. Some cover
types listed in the legend are not easily distinguished at the

scale of this map, including foothill shrublands on the lower


slopes of the mountains, small patches of irrigated land, and
the different kinds of riparian vegetation. The thicker dashed
lines show the location of two transects used for the profiles
of geological surface types and vegetation types, which are
shown in figs. 17.15 and 17.16. Cartography by Ramesh Sivanpillai, University of Wyoming.

The Laramie Basin

Fig. 17.8. (above) Mixed-grass prairie after a favorable growing


season in the northern part of the Laramie Basin. Common
grasses are blue grama, Indian ricegrass, junegrass, needleand-thread grass, Sandberg bluegrass, and western wheatgrass.
Elevation 7,100 feet.
Fig. 17.9. (left) True mountain-mahogany grows on the shallow soils of the ridges (Forelle Limestone), and black sagebrush dominates the shrubland in the foreground, underlain
by Satanka shale. The city of Laramie is in the background, at
an elevation of 7,165 feet.

Above the basin floor, where annual precipitation


is much greater, mountain forests are more common,
especially to the west, where dense forests of lodgepole pine, Engelmann spruce, subalpine fir, and occasional groves of aspen are widespread (see chapter 11).
Curiously, driving into the Medicine Bow Mountains,
one travels almost directly from the grasslands and
shrublands into lodgepole pine or aspen forest, unlike
elsewhere in the Rockies, where a widespread foothill
transition zone occurs, with juniper, limber pine, ponderosa pine, or Douglas-fir (see chapter 10). Much of the
west side of the basin floor seems too high for the usual
foothill transition. The Laramie Mountains on the
east are drier, and the typical foothill pattern is more

291

292 Landscapes of Special Interest


Fig. 17.10.Chimney Rock, also known
as Camel Rock, on the WyomingColorado border, southwest of Laramie.
The upper part is being carved from
Casper sandstone, the base from Fountain sandstone. Black sagebrush grows
on the sandy soil in the foreground.
This photo was taken in mid-November,
after the deciduous shrubs had lost
their leaves. The light gray shrub on
the gentle slope is true mountainmahogany, and the darker shrub on the
steep slope next to the base is skunkbush sumac. Elevation 7,700 feet.

apparent, with foothill shrublands of true mountain-

funnel moisture into places where roots concentrate.

mahogany and black sagebrush at lower elevations.

Both ponderosa pine and limber pine grow in such

At mid-elevations, on south slopes, scattered limber

microenvironments. Evidence indicates that the pines

pines and a few ponderosa pines grow with an under-

existed in such places about 6,000 years ago, when the

story of mountain-mahogany. On north-facing slopes

climate was warmer than it is today.19 However, then

at the same elevation, dense forests of limber pine and

as now, mixed-grass prairie grew over most of the

Douglas-fir are found, such as east of Laramie along

basin floor.

Interstate 80 passing through Telephone Canyon. Far-

Riparian woodlands, shrublands, meadows, and

ther up the canyon, forests or woodlands are found on

hayfields currently cover about 10 percent of the low-

both slopes. At the summit, such as near Lincoln Monu-

lands (figs. 17.12 and 17.13). Such a large proportion

ment, small stands of lodgepole pine, subalpine fir, and

of the landscape in riparian vegetation is unusual for

Engelmann spruce can be found on north slopes; limber

Wyoming and is attributable to flood irrigation (see

pine savannas occur elsewhere.

chapter 4).20 On the fringes of the riparian zone and

Trees rarely grow in the lowlands, except in riparian

in depressions, sometimes in association with saline

zones and in association with outcroppings of rock (fig.

shales, salts tend to accumulate on the soil surface,

17.11). Plausibly, the boulders and exposed soil provide

even to the point of forming a white crust (see chapters

favorable microenvironments for the establishment of

5 and 8). Only halophytes survive in such areas, includ-

tree seedlings, absorbing heat during the day that mod-

ing greasewood, saltbush, red swampfire, alkali saca-

erates cold temperatures at night. Also, there may be

ton, and saltgrass (about 4 percent of the basin floor)

less competition from other plants in such places, and

(fig. 17.14). Where water persists, small lakes develop,

the boulders and ridges may provide leeward environ-

with playa wetlands having bulrushes, spike


r ushes,

ments where snow drifting occurs, thereby providing

and in some places, cattails (see chapter 5). 21 Playa

more water. Moreover, when it rains, the water flows off

wetlands, marshes, and associated lakes and ponds

the rock surfaces into cracks or into the soil under the

are the core of three national wildlife refuges in the

rocks, where evaporation is minimal, which provides

basin, namely, Bamforth, Hutton Lake, and Mortenson

a longer-lasting source of water. Essentially, the rocks

Lake. Water diversions from Sand Creek to some of the

The Laramie Basin

lakes have been constructed to help maintain water


levels for the benefit of wildlife, especially migratory
shorebirds and waterfowl. However, the U.S. Fish and
Wildlife Service has junior water rights; thus, during
times of low streamflow, the refuges may not have
much water available. Mortenson Lake National Wildlife Refuge was established to protect the endangered
Wyoming toad (see fig. 1.9), known to exist only in the
Laramie Basin.
Elsewhere in the basin, the vegetation is more characteristic of sandy soils, such as south of Laramie in the
Red Buttes area and along Sand Creek. Small tracts of
dunes can be found, some stabilized by Indian ricegrass,
prairie sandreed, sandhill muhly, scurfpea, and silver
sagebrush (see chapter 9). Sand is supplied by the erosion of sandstones in the Casper Formation. Periodically, flash floods carry large amounts of sand from the
mountains.
Even more curious than sand dunes are the large
tracts of mima mounds in the basin, some on floodplains along the Big Laramie River and some in the
upland grasslands, such as southwest of Laramie (north
Fig. 17.11. Ponderosa pine and limber pine on an outcrop of
the Casper Formation on the east flank of Sheep Mountain.
Ponderosa pine occurs infrequently in the Laramie Basin, as
this species requires warmer summers with more rainfall than
has been characteristic for the basin.

of Harmony, and also near Hutton Lake National Wildlife Refuge, along the road to Chimney Rock; see figs.
9.9 and 9.10). The uncertain origin of these features is
discussed in chapter 9.

Fig. 17.12. Riparian woodlands east of Centennial


along the Little Laramie
River. The trees are narrowleaf
cottonwood. Grasslands and
shrublands extend into the
foothills of the Medicine Bow
Mountains. Limber pine is
common on the mountain
slope in the distance. Photo
by Ken Driese.

293

294 Landscapes of Special Interest

Fig. 17.13. (above) Riparian meadows and willow-dominated


shrublands along the Big Laramie River south of Woods Landing in Colorado. Note the lack of cottonwood in this area.
Also note the terraces and benches through which the river
flows. This topography was formed in the same way as along
the Snake River in Jackson Hole (see chapter 2).
Fig. 17.14. (left) Shrublands dominated by greasewood are
common in the Laramie Basin, occurring in low areas and
on the upland where the soils are saline and relatively moist.
Associated species include Gardners saltbush, pricklypear cactus, alkali sacaton, western wheatgrass, and Nuttalls alkali
grass. Greasewood often is found adjacent to playa wetlands,
as illustrated in fig. 5.6.

Geobotanical Relationships
Geobotany is the study of the relationships between
native plants and the underlying rocks on which they
grow. Such relationships have been considered in previous chapters, but several interesting patterns in the
Laramie Basin are noteworthy. One example is the
association of selenium-rich shales of the Niobrara Formation and a group of plants that only grow on soils
developed from this formation. Known as selenium
indicators, the group includes two-grooved milkvetch,
tineleaved milkvetch, woody aster, and princesplume. 22

The Laramie Basin

The Niobrara Formation in the Laramie Basin is exposed

by the availability of water next to the mountain, soil

on the west flank of Sheep Mountain and at scattered

chemistry over the Niobrara Shale and Forelle Limestone,

localities on the basin floor, including Big Hollow.

and a coarse soil texture (such as on the sandstones of

Though selenium is a requirement for animal health in

the Fountain, Casper, Muddy, and Cloverly formations).

low amounts, it is poisonous if too much is eaten.

Big sagebrush was found growing in relatively deep, well-

Another example is the distinctive way in which

drained soils derived from all rock types.

elongated aspen woodlands grow along the contour

Another informative approach to studying geo

on the east flank of Centennial Ridge, on the west side

botanical relationships is to examine the concomitant

of the basin (see fig. 10.3). In this case, the association

variation in topography, geologic surfaces, and land

is between aspen and surface exposure of the contact

cover along transects. Figures 17.15 and 17.16 illustrate

between permeable Paleozoic sedimentary rocks and

two profiles that were prepared for the Laramie Basin.

the impermeable Precambrian igneous and metamor-

One is 17 miles long and extends southeastward from

phic rocks that form the main part of the ridge. Water

the Little Laramie River to the Big Laramie River, pass-

flowing from above mostly stays near the surface, just

ing across the Airport Terrace and Big Hollow. The other

under the thin soils, until it meets permeable sedimen-

profile is 4 miles long, extending from the Big Laramie

tary strata of the Fountain Formation, where it infil-

River eastward to the Laramie Mountains, crossing a

trates the soil and bedrock. The whole mountainside

subdued ridge formed on the Chugwater Formation and

essentially funnels water to this contact zone, creating

a prominent ridge formed from the erosion-resistant

favorable conditions for aspen.

Forelle Limestone, across the erodible shales of the

In 1969, botanist William Myers studied the effects

Satanka Formation, and then up the lower outcrop of

of bedrock type on the abundance of native plants

the Casper Formation. The importance of shallow soils,

in the basin.23 With the assistance of noted geologist

salinity, and inundation from flooding can be seen. For

Brainerd Mears, he was able to identify nine distinctive

both transects, grassland plants are most common, sug-

sedimentary strata exposed on the west flank of Sheep

gesting that they tolerate a broad range of soil condi-

Mountain. Four of the strata were more resistant to ero-

tions and topography.

sion and were fully exposed as hogback ridges, whereas


others were more easily eroded and formed low, soilcovered troughs between the ridges. Some plants were

Reflecting on the Past, Anticipating the Future

found only on the Niobrara Formation, as might be

The Laramie Basin has experienced ecological changes

expected because of its high selenium content. They

throughout the Holocene and for millions of years be

included the selenium indicators mentioned previ-

fore that. Tectonic activity, fires, flooding, and climate

ously along with greasewood, saltbush, and winterfat.

change have been most influential. Recently, over

Other plants were found only on the resistant rocks of

shorter periods, the changes can be attributable to

the Fountain Formation, probably because they ben-

transitions in how native and EuroAmerican cultures

efited from a more favorable water supply (for example,

used the land (see chapter 2). Beaver were used hardly

aspen, shrubby cinquefoil, serviceberry, and common

at all by the Indians, but they were nearly extirpated

juniper). Another group of plants was found on the hog-

by EuroAmerican trappers (see chapter 2). The people

back ridges some distance from the mountain (currant,

who lived here for 10,000 years or more harvested only

mountain-mahogany, limber pine, and ponderosa pine).

a small number of trees, mostly small, in contrast to the

Myers found no plants that grew only in the troughs

immigrants arriving about 150 years ago. They needed

between the ridges, probably because these low areas

railroad ties and lumber for construction. Unfenced

had soils that were influenced as much by weathering

bison herds roamed the plains until the 1860s, when

of the adjacent ridge rocks as by the underlying shales.

cattle and sheep replaced them. Large numbers of

Myerss study illustrated that rather few species of

domestic livestock survived on the prairie vegetation.

plants are restricted to one geologic formation. In his

Sheet-wash erosion left some slopes with little or no

study area, plant distribution is determined primarily

soil, most likely because of heavy livestock grazing near

295

296 Landscapes of Special Interest

Elevation, feet

Modern Little
Laramie River
floodplain
7,500

Little
Laramie
River
terrace

Pleistocene
floodplain
Airport Bench

Eolian deflation Dissected Laramie River


basinBig Hollow bench
terrace

Modern
Laramie
River
floodplain

Dissected
bench with
Pleistocene
overwash

7,400
7,300
7,200
7,100
7,000

Northwest
0

10,000

Southeast
20,000

30,000

40,000

50,000

60,000

70,000

80,000

90,000

Distance along profile, feet


Riparian
Upland
grassland,
hayland

Upland
grassland

Fig. 17.15. Seventeen-mile profile across the Laramie Basin


with topographic features and land cover indicated. As
shown in figs. 17.1, 17.3, and 17.7, this long profile extends
southeastward from the Little Laramie River, where a mosaic
of riparian woodlands, shrublands, meadows, and irrigated
hayland can be found. Grasslands are found over most of the
upland, across the Airport Bench, and down the slopes to near
the bottom of Big Hollow, where a playa with an ephemeral
pond has formed. There is no drainage from Big Hollow, but
standing water is rare because of low precipitation and no
streamflow. The Airport Bench is a remnant of the valley
floor at an earlier phase in the excavation of the basin and
has a surface that has been armored by coarse river cobbles

Not
vegetated,
playa Upland
grassland

Irrigated Upland
Upland
hayland grassland
grassland
Riparian

deposited during the Pleistocene. From the Big Hollow playa


southeastward, grasslands are found on the slopes and top
of a dissected bench. Beyond that, the former floodplain of
the Big Laramie River, known as a terrace, has a mosaic of
grasslands, shrublands, and irrigated hayland. Several lakes
also are located in this area.Riparian vegetation is found in
the current floodplain of the Big Laramie River. The profile
terminates on the dissected margins of a third, more ancient,
erosion-deposition surface. The Harmony Bench, illustrated in
fig. 17.5, does not occur along this profile. Plant species composition is influenced by topography, varying soil characteristics, and exposure to wind.

sources of water and railroad stockyards (fig. 17.17).

Irrigation had two general effects. First, it disrupted

Wolves were eradicated, enabling coyotes to become

the streamflow patterns of the rivers and creeks, reduc-

more abundant. Prairie dog densities declined; down-

ing flood peaks in the spring and emptying many

cutting of stream channels accelerated.

streams by late summer, when irrigation water was in

Equally profound change resulted from the establish-

greatest demand. Second, it enlarged the area of land

ment of irrigation systems, beginning in the 1870s.24 The

that was flooded. Essentially all floodplains and some

broad floodplains of the Big Laramie and Little Laramie

adjacent uplands were altered by this practice. In gen-

rivers enabled flood irrigation (see figs. 17.1 and 17.7).

eral, native riparian woodlands and shrublands exist

Dams, canals, and ditches eventually were developed,

today only on land that was too wet for domestic use,

including the Wheatland reservoirs on the Laramie

or where the presence of native plants provides benefits

River. Some of this irrigated land, though a small pro-

appreciated by the landowners, such as shade for live-

portion of the basin floor, was originally tilled for crops,

stock, erosion control, shelter for homes and barns, or

eliminating the native vegetation, but this practice soon

habitat for fish and wildlife.

proved uneconomical because of the cool, short grow-

Today the glint of numerous lakes across the basin is

ing season. Livestock production prospered; irrigated

especially obvious at sunset from the crest of the Lara-

haylandnow with mostly introduced speciesbecame

mie Mountainsall more or less permanent because of

a valuable commodity.

their connection to irrigation systems and providing

The Laramie Basin

Big Laramie
River floodplain

Chugwater
Formation

Tertiary
alluvium

Chugwater
Formation

Forelle
Limestone

Satanka
Shale

297

Casper
Formation

Elevation, feet

7,600
7,400
7,200
7,000

West

East
5,000

10,000

15,000

20,000

Distance along profile, feet


Riparian
meadow,
shrubland

Meadow,
upland
grassland

Not
Riparian
vegetated, meadow
landfill

Upland
grassland,
sagebrush

Riparian
Sagebrush,
meadow
mtn.-mahogany
Foothill
Upland
shrubland
grassland,
sagebrush

Fig. 17.16. Four-mile profile just north of Laramie, from the


Big Laramie River into the foothills of the Laramie Mountains, illustrating topographic features and land cover (see
figs. 17.1, 17.3, and 17.7 for location). Toward the east, beyond
the riparian meadows and shrublands, the surface slopes up
onto red beds of the Chugwater Formation. This rock is more
resistant to erosion than most materials in the middle of the
Laramie Basin, but the sloping surface has been altered by
erosional processes. At about 8,000 feet along the profile,
alluvial materials from higher slopes cover the Chugwater.
Soil development throughout this area is sufficient to support
grasslands. The portion of the profile lacking vegetation is the
Laramie Landfill. A narrow riparian meadow is intercepted
just beyond the landfill, where the profile crosses the bed of
an ephemeral stream. The steeper Forelle Limestone then

rises to a more pronounced hogback ridge with shallow soils


supporting foothill shrublands. The ridge is sharply defined
because of erosion of the underlying Satanka Shale beds to
the east. The profile ends on the slope of the Casper Formation, which extends to the crest of the Laramie Rangethe
hydrologic boundary of the basin. The lowest outcrops of the
Casper Formation mark the boundary of the valley floor for
much of the eastern side of the basin. Grasslands are characteristic of deeper soils developed over the Chugwater and
Satanka formations; shrublands with mountain-mahogany,
Rocky Mountain juniper, black sagebrush, and wild currant
are found on the ridges of the Forelle and Casper formations.
The Casper Formation is a very important aquifer, as discussed in the text, and its limestone strata are an important
source of calcium carbonate for the local cement plant.

wetlands that are important for wildlife (see chapter

endangered speciesthe whooping crane, piping plo-

5).25 Overall, irrigation has retained some of the water

ver, least tern, and pallid sturgeon. The Nebraska species

that previously flowed out of the basin, spreading water

depend on adequate water in the Platte River, of which

across the lower, flatter parts of the landscape, thereby

the Laramie River is a tributary. Potentially, more irri-

producing better-watered ecosystems than had existed

gation upstream in Wyoming, or more severe droughts

previously. Water withdrawals modified the rivers, but

anywhere in the watershed, could reduce the amount

irrigation waters that contributed to the formation of

of water downstream. For whatever reason, if more river

wetlands appear to have increased biological diversity

water is required to maintain viable populations of the

in the basin as a whole. Introduced trout still thrive in

endangered species in Nebraska, the Platte River Com-

the rivers, as they do in the lakes that had been created

pact and other regulatory statutes could require more

or enlarged. As noted, Congress designated some associ-

efficient use of Wyoming and Colorado water.27 That

ated wetlands as national wildlife refuges. Some species

could reduce water availability for Wyoming wetlands.

clearly benefited from wetlands maintained with irriga-

Can irrigation systems be engineered to use water more

tion projects, including mosquitoes. 26

efficiently and so enable survival of the Nebraska endan-

A conundrum has now developed that connects Wyo-

gered species without leading to the decline of Wyo-

ming wetland habitatand the endangered Wyoming

mings endangered species? That challenge would be

toadwith wetlands in Nebraska, where there are four

especially difficult if climate change leads to extended

298 Landscapes of Special Interest

Fig. 17.17.Exposed limestone of the Forelle Formation on the


east edge of Laramie now supports a shrubland of mountainmahogany. Most likely, parts of this area were mixed-grass
prairie prior to the arrival of the Union Pacific Railroad in
1868, when large numbers of cattle and sheep congregated
in the area because of close proximity to the city springs and
a spur of the railroad, where the animals were loaded onto
stock cars. Erosion over most the area would have been an

ongoing process because of the slope of the underlying bedrock, but it probably was accelerated by heavy grazing. Flash
floods are known to occur in the area, which would have
contributed further to soil loss. A remnant of the prairie can
be seen in the central part of the photo, where the soil has
not yet eroded. Limestone was quarried in this area, first for
building construction and later for the manufacture of plaster
and cement.

droughts. The way current laws are written, water rights

the potential for more wind farms, transmission lines,

for wildlife benefits have low priority.28

and roads (see chapter 18).

Another kind of change is associated with rural sub-

As in ecosystems throughout the region, invasive

divisions, especially around Laramie, along rivers, and

plants are a growing problem. 29 The grasslands and

in the foothills and mountains. As people move to the

shrublands of the Laramie Basin are still relatively

country, former grazing land is converted to residen-

free of invasive plants, though cheatgrass is becoming

tial use, wildlife habitat is fragmented, open space is

more common in the foothills. The riparian zones are

diminished, and the potential for problems with inva-

another matter, where Canada thistle, Dalmatian toad-

sive plants increases (see chapter 18). Moreover, the

flax, perennial pepperweed, white top, and spotted and

potential pollution of the Casper aquifer by a few land-

Russian knapweed are common. Some of their effects

owners on the outskirts of Laramie is a concern for the

are discussed in previous chapters. Native meadow

entire community. Guidelines have been proposed that

grasses that originally were harvested for hay included

would protect the aquifer, but this necessarily involves

meadow brome, redtop, and tufted hairgrass, but with

the cooperation of landowners who have other visions

irrigation they have been largely displaced by intro-

for their land. Currently, rural subdivisions occupy

duced species, such as foxtail, redtop, smooth brome,

only about 8 percent of the basin floor, but much of

and timothy. 30 One variety of foxtail, Garrison creeping

that is concentrated in the foothills, and the land area

foxtail, was introduced unintentionally when hay bales

involved increases annually as more people seek alter-

were brought to the basin from elsewhere. Initially,

natives to living in town. Accentuating such trends is

some ranchers appreciated its forage value, but now it

The Laramie Basin

43

6.0

42

5.5

41

5.0

40

4.5
4.0

39

3.5

38
1890

1910

1930

1950

Year

1970

1990

2010

Mean Annual Temperature (C)

Mean Annual Temperature (F)

6.5

3.0

Fig. 17.18. Mean annual temperature for Laramie from 1895


to 2011. An increasing trend since about 1982 is suggested,
though the highs and lows are still within the range of historic
variability. The solid horizontal line is the average temperature for the period of record (40.3F); the dashed lines are 1

standard deviation. See also fig. 3.9. From the Wyoming Water
Resources Data System, prepared with PRISM Climate Group
data, Oregon State University (http://www.prism.oregonstate
.edu/terms.phtml), courtesy of Christopher Nicholson.

has displaced many other meadow species. The hay,

estimate of trends, meteorologists now combine infor-

though abundant, is generally considered to be inferior

mation about typical storm patterns with data from a

to that existing previously.

Restoration to more desir-

network of weather stations to calculate annual precip-

able species will be difficult, should that be considered

itation for tracts of ground roughly 2.5 miles square.

a practical and worthwhile goal.

When this is done for Laramie, a slight downward trend

31

Climate change is likely to influence basin eco

is apparent (fig. 17.19). The mean annual precipitation

systems even more than other human activities. Chap-

in the past 10 years is 9 percent less than in the previous

ter 3 includes a discussion of climate change in general,

decade (11.1 inches instead of 12.2 inches).

and the potential effects are considered throughout

Similarly, climatologists have integrated precipita-

this book. For the Laramie Basin specifically, even

tion data for large drainage basins. Specifically, scien-

though mean annual temperatures during the past 10

tists with the National Climate Data Center divided

years are still within the range of historical variabil-

the state into ten basins and produced historical mean

ity, the record suggests a gradual warming trend since

annual precipitation for each year going back to 1895. 32

the mid-1980s (fig. 17.18; see also fig. 3.9). With regard

A cursory examination of the ten graphs suggests that

to precipitation, the trends are more difficult to ascer-

six of the basins have had no obvious trend upward or

tain. However, unless annual precipitation increases

downward during a period of 116 years. Four, includ-

along with temperature, a warmer climate will lead to

ing the Upper Platte River Basinof which the Lara-

more rapid drying and more frequent water stress dur-

mie Basin is a parthave had slight downward trends

ing the growing season.

(see fig. 17.19). Significantly, none of the basins had an

Though inconclusive, a closer look at precipitation

upward trend over the entire period of record. With

records is warranted. The most intuitive approach is to

warming, climate scientists predict that there will

look at records from a single weather station. However,

be less snowfall and more rainfall during the winter

there is great variation from year to year. Also, precipi-

months, and that snow cover will disappear earlier

tation events tend to be patchy across the landscape;

in the spring. As described in chapter 3, this trend is

downpours occur in one area but not nearby. For a better

already happening. Even if annual precipitation stays

299

19

Laramie

Annual precipitation (Oct-Sept)


5 year average precipitation
1895-2012 average precipitation

45

40
15
35
13
30
11

Annual Precipitation (cm)

Annual Precipitation (inches)

17

25
9
20
7
19

Upper Platte River Basin

Annual precipitation (Oct-Sept)


5 year average precipitation
1895-2012 average precipitation

45

40
15
35
13
30
11

Annual Precipitation (cm)

Annual Precipitation (inches)

17

25
9
20
7
1900

1910

1920

1930

1940

1950

1960

1970

1980

1990

2000

2010

Year (October - September)


Fig. 17.19.The top graph shows annual precipitation from
1895 to 2012 for a 2.5-square-mile tract of land in the vicinity
of Laramie, based on a network of actual data and knowledge
about snow and rainfall events in the area. The bottom
graph shows the same for the Upper Platte River Basin, which
includes the Saratoga Valley and Laramie Basin. A downward trend is suggested. The dashed line in both graphs is 1
standard deviation from the average for the period of record
(11.8 inches). The red line is a 5-year running average. Average
annual precipitation for the period of record is 12.1 inches.

The average annual precipitation in the early 1900s, when


irrigation systems and Wyoming water law were developed,
was 34 inches higher than at present. The top graph is from
the Wyoming Water Resources Data System, prepared with
the PRISM Climate Group data, Oregon State University,
http://prism.oregonstate.edu. The bottom graph is from the
Wyoming Water Resources Data System, prepared in collaboration with the National Climatic Data Center (http://
www.ncdc.noaa.gov/oa/ncdc.html), courtesy of Christopher
Nicholson.

The Laramie Basin

about the same, drought stress is likely to be more fre-

increase in mean annual temperature of 25F, which

quenta trend that is apparent for the western states

would likely increase potential evapotranspiration by

as a whole (see fig. 2.11).

0.010.04 inch/day. For this exercise, the first scenario

Overall, the data suggest a slight decline in annual

proposes increased winter precipitation and decreased

precipitation in the Laramie Basin during the past

growing season precipitation. The net effect would

10 years, a trend further substantiated by declining

cause a shift from grassland to mixed desert shrubland.

streamflow in some rivers, as is illustrated for the Green

The second scenario projects the same increase in tem-

River where it flows into Flaming Gorge Reservoir.33

perature, but with increased summer precipitation

However, no similar decline occurred for the Laramie

rather than winter precipitation. This scenario would

River. There is the potential that summer precipitation

favor continuing dominance of grassland, but with an

might be more closely associated with plant growth in

increase in cheatgrass. Ponderosa pine may also become

grasslands than is annual precipitation, but such analy-

more widespread than it is today, because of its apparent

ses have not yet been done for the Laramie Basin. Also,

reliance on warmer, wetter summers. 35

winter precipitation can be a significant factor for grasslands if it results in considerable water in the soil at the

In general, much remains to be learned about the impli-

end of winter.

cations of climate change combined with habitat frag-

Two scenarios for climate change have been proposed

mentation, invasive plants, and the long-term effects

for the east-central Rocky Mountain region that are per-

of living on the landa topic discussed further in the

tinent for the Laramie Basin.

next chapter.

34

Both assume a plausible

301

This page intentionally left blank

Part Six Sustainable Land

Management

This page intentionally left blank

Using Western
Landscapes
Chapter 18

The first EuroAmerican immigrants to Wyoming Terri-

West Coast along the Oregon Trail. 2 A few of them estab-

tory began building their homes in the 1820s. The human

lished homes along the way, but the first large wave of

population at that time numbered fewer than 10,000, 99

immigrants arrived after the completion of the Union

percent of them tribe members.1 Most likely, bison, elk,

Pacific Railroad in 1869. Others arrived in the 1950s and

deer, and pronghorn outnumbered people by 150 to 1.

1970s, when Wyomings oil, coal, uranium, natural gas,

Then, and for the previous 10,000 years or more, the

and soda ash became more valuable to the nation as a

primary human uses of the land were hunting and gath-

whole. Upticks in the annual census occurred with each

ering. Forty years later the population of domesticated

new development and were mostly associated with the

livestock still numbered only a few thousandmostly

energy industry. Tourists, campers, hunters, and anglers

horsesand timber harvesting was restricted to small

come to Wyoming during the summer and fall.

trees used for fuel, shelters, sleds, and weapons. The great-

During the past 200 years, new episodes of rapid

est human influence was igniting fires that sometimes

change have occurred. The world of Native Americans

burned for weeks or months. Tribal council meetings to

was turned upside down by the slaughter of bison and

discuss air and water quality were probably not necessary.

the introduction of horses, cattle, and sheep. Nothing

Today there are about 575,000 people and somewhat

so severe has happened since then. Still, the rate of

more than a million elk, deer, bison, moose, and prong-

population growth in western states has been above the

horn in Wyoming. Livestock, mostly cattle and sheep,

national average. Wyomings population grew 14 per-

number almost 2 million. A small proportion of the

cent from 2000 to 2010, with some counties growing

land is cultivated; agriculture is concentrated in a few

much more rapidly. For example, Campbell Countys

areas at low elevations. Fossil fuels rather than wood

population in the Powder River Basin grew by 37 per-

are the primary source of energy. Agriculture, techno-

cent during that 10-year period, and Sublette County in

logical advances, and towns and cities have benefited

the upper Green River Basin grew by 73 percent.

many people, but they have also led to a need for envi-

New oil and natural gas developments were the pri-

ronmental quality councils at federal, state, and local

mary cause for this latest wave of immigration, with

levels. Conservation biologists and land managers are

thousands of new wells drilled in the Powder River Basin

faced with new challengesthe topic of this chapter.

alone. Such activity has threatened biological diversity


through habitat fragmentation. An added threat has been

The New West

new rural housing developments, which provide places


where people can live close to the natural amenities that

Hundreds of thousands of EuroAmericans passed

more and more people enjoygood hunting, fishing,

through Wyoming in the mid-1800s, destined for the

hiking, and biking. But traffic has increased along the


305

306 Sustainable Land Management

Fig. 18.1.Open spaces have persisted largely because of short,


cool growing seasons and the challenges associated with
obtaining water. However, new roads and homes are now

being constructed in such places on private land, a trend that


is characteristic of the New West. The implications can be
significant. Photo by Nick Fuzessery.

thousands of miles of new roads that are required, and

of the past 20 years have had noticeable adverse effects

invasive plants have expanded their ranges. Among the

on many highly valued resourcesattracting the atten-

consequences are dramatic declines in the abundance of

tion of state governors and members of Congress. To

some native species, especially various birds found pri-

illustrate, Governor Jim Geringer sponsored work-

marily in shrublands and grasslands.

ing groups on the protection of open space, Governor

Alarmed by such trends in what some people refer

Dave Freudenthal promoted the sage-grouse conserva-

to as the New West, Chris Madson, long-time editor of

tion initiative and the Wyoming Wildlife and Natural

Wyoming Wildlife, wrote in 2007:

Resources Trust (approved by the Legislature in 2005),

Its not that we havent been working on conservation


in the last thirty years. Weve cleaned up our air and
water; weve pulled species like the black-footed ferret
and whooping crane back from the brink, but overall,
the conservation effort of the last generation is clearly
not up to the challenge of offsetting the damage to
land and wildlife we continue to inflict. . . . What will
Wyomings streams look like in twenty years, when
the last of the high glaciers in the Wind River Range
fades away to nothing? Can we adequately mitigate
the impacts of the latest round of gas drilling if the
current ten-year drought in sagebrush country continues for thirty, fifty, even a hundred years?3

and Wyomings congressional delegation sponsored


the Wyoming Range Legacy Act, which withdrew some
lands from future oil and gas leasing. The current governor, Matt Mead, supports the sage-grouse conservation
initiative and is promoting policies that may mitigate
some of the adverse effects of energy development.
Thanks to such initiatives, some concerns have been
addressed by state laws. Its clear that many residents
place great value on, for example, land reclamation
after mining, the protection of air and water quality,
the proper location of industrial activities, and the wise
management of threatened but still-abundant fish and
wildlife.4 Moreover, support is growing for the improved
management of sensitive areas, such as riparian zones

In 1994, when Mountains and Plains was first pub-

and other habitats critical for threatened species, and

lished, the generally accepted sentiment was that over-

for the curtailment of the spread of invasive plants and

all, Wyoming landscapes were about as pristine as could

animals. Such problems are discussed in previous chap-

be found in the lower forty-eight states (figs. 18.1 and

ters. Here the focus is on water development, multiple

18.2). That may still be true. However, the rapid changes

use on public lands, habitat fragmentation, the concepts

Using Western Landscapes

Fig. 18.2.Cultivated land, usually restricted to lowlands


where water is accessible, provides open space as well. The impact on the native flora and fauna is significant, but the land
still provides habitat for some kinds of wildlife, and many

people appreciate pastoral vistas, such as this one. In the New


West, land trusts work with property owners interested in
conserving the habitat and ecosystem services provided by
such places.

of ecosystem management and ecosystem services, and

onto the land, but only at great cost. New reservoirs and

adaptation to climate change.

trans-basin diversions are on the planning boards, and


existing diversions are being upgraded. State laws and reg-

Water Development

ulations have placed constraints on such developments


reflecting the view of many that the few unregulated

Not surprisingly, water is a limiting factor for human

rivers that remain are a valuable resource. Also, though

activities throughout the region. The traditional solu-

still difficult to implement, state laws now recognize

tions have been to construct reservoirs, irrigation sys-

instream flows as a beneficial use of water, and portions

tems, and trans-basin diversions, thereby making water

of some rivers are now protected under the Wild and Sce-

available wherever economically feasible and politically

nic Rivers Act of 1968. In Wyoming, the protected rivers

desirable (fig. 18.3). Faced with interstate competition

are the headwaters of the Snake River and the Clarks Fork

for water and the rights associated with prior use, a

of the Yellowstone River. Combined, they represent less

common attitude has been to use it or lose it. Pro-

than 0.4 percent of the states river miles.5

moted historically by the U.S. Army Corps of Engineers,

Groundwater is another issue. The Ogallala (High

the U.S. Bureau of Reclamation, and local chambers of

Plains) aquifer on the western Great Plains, extending

commerce, and heavily subsidized by state and federal

from Nebraska and eastern Wyoming into New Mex-

grants and loans, such projects have indeed stimulated

ico and Texas, has supplied agriculture and munici-

the economy. Also, the projects have often created

palities for many years, but the rate of withdrawal

additional benefits, such as flood control, hydroelectric

exceeds the rate of recharge. Significant drawdowns

power, and new recreational opportunities.

of this aquifer have forced some farmers to revert to

Still, regulated rivers have replaced rivers that flowed

dryland farming.6 Furthermore, groundwater in some

freely long enough to cut canyons through mountain

areas has been degraded through the deep percolation

ranges, and many reservoirs will probably be silt laden

of polluted surface water. Clearly, humans have the

in less than a century. Reservoir silt can be pumped back

capacity to affect deep groundwater supplies. Recent

307

ACRE-FEET (millions)

308 Sustainable Land Management


15
14
13
12
11
10
9
8
7
6
5
4
3
2
1
0
1890

to develop more efficient irrigation systems. Indeed,


people from all walks of life are striving to reduce the
amount of water they use.

Multiple Use on Public Lands


For many years, little attention was given to land management activities on federal lands administered by
the Forest Service and the Bureau of Land Management. However, by the late 1950s, concerns were being
expressed about excessive timber harvesting and live1910

1930

1950

YEAR

1970

1990

2010

Fig. 18.3. Water storage capacity in Wyoming reservoirs has


increased very little during the past 50 years. The last major
project was the High Savery Reservoir on the Little Snake
River, completed in 2004 and located northwest of Baggs.
Though locally important, it added only 22,433 acre-feet of
storage capacity. One acre-foot is the volume of water required
to cover one acre of land with one foot of water. Based on data
from Ostresh et al. (1990) and recent public records.

stock grazing. Responding to public sentiment in 1960,


Congress passed the Multiple UseSustained Yield Act,
a law that emphasized the importance of recreation,
watershed, and fish and wildlife values on national
forest lands as well as that of timber, livestock forage,
and other commodities. A similar philosophy was eventually adopted by the Bureau of Land Management,
though Congress did not formalize it until passage of
the Federal Land Policy and Management Act in 1976.
Many are comfortable with the multiple use concept.

concerns have emphasized the potential impact of

Timber harvesting and the extraction of minerals and

toxic hydraulic fracturing fluids, used in a process

fossil fuels are still possible, but public lands are now val-

known as fracking to extract oil and gas that previ-

ued for more than these resources alone. At times, some

ously was inaccessible.

stakeholders believed that the mandates of the Multiple

The demand for water has led to the consideration

UseSustained Yield Act were being abandoned, such as

of cloud seeding with silver iodide and other nucleat-

when the Wilderness Act of 1964 was passed. However,

ing agents as a means of increasing precipitation. Some

as the 1960 law states, multiple use is not necessarily

observations suggested that seeding might increase rain

the combination of uses that will give the greatest dol-

and snowfall by 15 percent or more, at least during some

lar return or the greatest unit output. Multiple use does

years. If droughts caused by climate change become

not require resource extraction everywhere.

more common, augmenting water supplies in any way

Much of the current debate over public land manage

is attractive. To be effective, however, the seeding must

ment centers on the loss of wildlands, whether from oil

happen when moisture-laden clouds exist, which are

and gas development, mining, or timber harvesting, and

less common during drought. Thus, at the time extra

the roads associated with these activities (fig. 18.4). Tim-

water is most needed, the likelihood of producing more

ber harvesting was the primary concern 20 years ago,

rain or snowfall is low. A committee of scientists orga-

but now controversy over fossil fuel extraction is in the

nized in 2003 by the National Research Council, an arm

news more often. Oil and gas developments are promoted

of the National Academy of Sciences, concluded there is

because of their high economic value and the potential

no convincing evidence that cloud seeding is effective.7

for less dependency on other nations; timber harvest-

This committee also doubted that the cost-to-benefit

ing is promoted because it, too, contributes to the local

ratio would be low enough to justify such a herculean

economies, and sometimes tree cutting can be used to

effort to increase water supplies.

improve wildlife habitat (see chapters 11 and 12). How-

Desires for more water are easily rationalized, but

ever, others maintain that the remnants of wildlands

learning to live with the amount provided by natural

now cover a very small area and that the best use of such

weather and hydrologic patterns is likely more sus-

areas is for wildlife habitat, dispersed recreation, and

tainable. In this spirit, agriculturalists are working

solitudeamenities that attract many people to the area.

Using Western Landscapes

Fig. 18.4. Habitat fragmentation from timber harvesting is


clearly visible around the mostly forested Savage Run Wilderness, located in the Savage Run Creek watershed on the west
slope of the Medicine Bow Mountains. This is the only entire
watershed in the Wyoming portion of this mountain range

The role of the public in public land management has


been evolving over recent decades. In the early days of the

without roads. Red indicates forest meadows; areas of gray on


the left are foothill shrublands. The geometric shapes in the
forests are clearcuts done at various times prior to the 1980s.
Landsat Thematic Mapper image, June 24, 1991. Cartography
by Ken Driese.

public interactions. Multiple use is better understood; the


values of wildlands are more widely appreciated.

U.S. Forest Service, for example, management decisions


lay almost exclusively in the hands of the agencys foresters. After World War II the public took more interest in

Habitat Fragmentation

what was happening on their lands, and such legislation

Long before Europeans arrived in North America, grass-

as the 1969 National Environmental Policy Act mandated

lands and shrublands extended as far as the eye could

that the Forest Service inform people of significant man-

see. In one direction or another, however, there were

agement issues and proposed actions. During the past

ridges, buttes, riparian zones, sandhills, or mountain

20 years, partnerships have sprung up across the West,

rangesislands on the plains. With variety in available

as illustrated by the Uncompahgre Plateau Partnership

environments, a greater diversity of plants and animals

in western Colorado. Community members from diverse

could survive. Bison surely found shelter on the lee sides

backgrounds join Forest Service and Bureau of Land

of ridges, as did Native Americans.8 The landscape was

Management employees in evaluating forest conditions

patchy, but at a rather coarse scale.

and monitoring the effects of forest management activi-

Today, aerial views show a more intricate patchwork

ties. This collaboration greatly increases understanding of

of croplands, human shelters, groves of trees, oil and

the complexities of land management while building trust

gas fields, and wind energy projectsall connected by

between the agencies and the public. Participants include

roads and superimposed on the landscapes that were

representatives of timber, grazing, motorized recreation,

there before. Many populations of animals have found

and environmental interests. As a result, forest manage-

their habitats broken into patches too small for their

ment has moved forward with less of the conflict and

survival, or too isolated from other favorable patches to

polarization that had previously characterized agency-

enable migration, reproduction, and the procurement of

309

310 Sustainable Land Management


Fig. 18.5. Most of Wyoming is accessible by roads, though many are not
easily traveled or are private. With or
without fences, they affect the movements of pronghorn, deer, elk, and
sage-grouse. Moreover, their impacts
extend beyond the roadbed, contributing further to habitat fragmentation.
Photo by Ken Driese.

food. Some native animals have become locally extinct,

To illustrate, elk and mule deer in shrublands tend

or their populations have declined. Examples in forested

to stay about 1.5 miles away from roads in summer,

landscapes include the lynx, wolverine, grizzly bear, great

when traffic is greatest, and about 0.75 miles away

gray owl, boreal owl, northern goshawk, and northern fly-

from roads in winter.9 Much of their time is spent in

ing squirrel. In grasslands and shrublands, declining spe-

roadless areas. Similarly, sage sparrows and Brewers

cies include the sage-grouse, sage sparrow, sage thrasher,

sparrows tend to avoid roads and oil wells by about

Brewers sparrow, dickcissel, burrowing owl, lark bunting,

300 feet, and sage-grouse by 23 miles. Roads also are

bobolink, ferruginous hawk, prairie falcon, upland sand-

pathways along which invasive plants become estab-

piper, McGowns longspur, chestnut-collared longspur,

lished and expand.

Columbian sharp-tailed grouse, long-billed curlew, swift


fox, prairie dog, black-footed ferret, and pygmy rabbit.

Other effects of roads include direct mortality from


collisions. The number of animals killed by vehicles is

The adverse effects of fragmentation are most severe

small compared to fatalities from inadequate food sup-

during the inevitable episodes that lead to other stresses,

ply and hunting, but large ungulates on the highway

such as drought, severe winters, high levels of predation,

are genuine traffic hazards for people. Also, crossing

and avoidance behavior when humans or predators are

busy highways is one added stress that diminishes the

present. Even fires, beneficial for maintaining biological

energy in body fat required for surviving the winter. If

diversity since prehistoric times, can be detrimental if

energy reserves are insufficient, the animals are unable

they further fragment an already marginal habitat. Habi-

to return to their summer range, where their reserves

tat fragmentation can push animals into places where

are normally restored (see chapter 15).

they encounter excessive competition from others already

The best-known example of road effects on ungu-

there, or to habitats that are marginal, which reduces the

lates is at Trappers Point, along U.S. Highway 191 west

likelihood of survival during inevitable difficult times.

of Pinedale. This highway intersects the fall migra-

Roads, rural subdivisions, timber harvesting, and

tion route of pronghorn from Jackson Hole and Grand

energy development are common causes of habitat frag-

Teton National Park up the Gros Ventre River valley,

mentation and warrant further discussion (fig. 18.5).

across the divide to the headwaters of the Green River,

Roads

and then southward along the Green River valley to


the sagebrush steppe south of Pinedale and east of Big

The very presence of roads reduces the amount of hab-

Pineythe longest known migration of a mammal

itat far more than the area required for the roadway.

in the contiguous United States. If they survive, the

Using Western Landscapes

Fig. 18.6. Forest fragmentation by clearcutting in the


Medicine Bow Mountains. This mountain range is relatively
flat and accessible, which facilitates harvesting over larger
areas than, for example, in the more rugged Wind River and
Bighorn mountains. This aerial photo was taken in the 1980s.
New trees (mostly lodgepole pine) are now growing in all the
openings. Current management prescriptions call for smaller,

irregularly shaped openings, and tree thinning has become


more common than clearcutting. Much of the harvesting in
recent years has been of beetle-killed trees, which provides a
source of wood while reducing hazardous snags along roads
and near campgrounds. Another motivation for harvesting
in some areas is to improve the habitat for specific kinds of
wildlife. See also fig. 18.4. Photo by William K. Smith.

pronghorn return to Jackson Hole in the spring along

servation easements, bear-proof containers, and habitat

the same route. Collision mortality and the costs of

enhancements elsewhere.11 The effectiveness of such

vehicle repair have been a serious concern for many

measures is controversial.

years. To solve this problem, in 2012 the Wyoming


Highway Department completed a system of 25 miles
of 8-foot-high fences that funnel the pronghorn and

Timber Harvesting and Energy Development

deer through six underpasses and across two over-

The discussion of habitat fragmentation began about

passes along a 13-mile stretch of road. The $10 million

30 years ago with evaluations of the silvicultural prac-

cost of the project was justified largely by less vehicle

tice known as clearcutting, whereby all trees are cut

damage. Placing a value on each animal killed is still

whether they are harvested or not (fig. 18.6; see chap-

difficult.10 Without accommodations for wildlife, such

ter 12). The rationale for clearcutting was to enhance

as at Trappers Point, widening and dividing highways

the growth and health of new lodgepole pine trees by

to enable faster, safer travel for motorists has the oppo-

providing them with the maximum amount of light,

site effect on animals.

water, and nutrients, thereby mimicking some effects

New roads are often constructed into the foothills,

of a crown fire. Predictably, logging roads fragmented

because homebuilders are attracted to scenic locations

the forest still further, with the impact extending well

adjacent to national forests and national parks. Un

beyond the roadway. Conservation biologists distin-

fortunately, such places also are important habitat for

guish interior from edge forests, and have identified

some kinds of wildlife, especially during the winter.

species, mostly animals, that require one or the other.

People who live there commonly appreciate the abun-

Timber harvesting and road building can reduce the

dance of these animals, and sometimes they attempt

amount of interior forest and increase the amount of

to mitigate their own adverse effects, such as with con-

edge forest.

311

312 Sustainable Land Management

Debates often centered on whether natural forest disturbances, such as fire and blowdowns, had the same
effect as tree cutting (see chapter 12). Commonly, the
landscape patterns caused by natural disturbances in
wildland landscapes are thought to enhance biological
diversity, not because all native organisms find them
beneficial, but rather because the population sizes of
some species drop while others rise after the disturbance. The diversity exists because of the changes in the
landscape over long periods of time. Some landscapes
are large enough to accommodate such disturbances
without the adverse effects of fragmentation. Others are
already so patchy that further timber harvesting would
most likely lead to the loss of some species.
Although timber harvesting and roads have surely
contributed to forest fragmentation in some areas (see
fig. 18.4), some forests are naturally divided into small
patches, that is, mostly edge forest, because of abrupt
differences in mountain topography and soils.12 Harvesting might further fragment the habitat in such
places, but there is reason to question whether interior
forest is broken up in the process. The loss of interior
forest may be most significant in places of relatively flat
topography and extensive forest, such as in some parts
of the Medicine Bow Mountains.

Fig. 18.7. Habitat and open space fragmentation by oil and


gas development is now common in parts of the Powder River,
Green River, and Washakie basins. Photo Dave Showalter,
Lighthawk Aerial Support.

With regard to energy projects, a few well pads and


mines scattered about the landscape would have relatively little effect on wildlife. However, in the Powder

tors that use the towers for perches, such as ferruginous

River Basin alone, the energy boom during the past

hawks and golden eagles. All kinds of energy develop-

20 years has led to more than 10,000 new wells for

ment fragment wildlife habitat and have impacts on an

oil and gasall serviced by several thousand miles of

area larger than the land they occupy.13

new roads (fig. 18.7) and having adverse effects on air


quality. Similar intensities of development are occurring in other parts of the state, such as in the Green

Ecosystem Management

River and Washakie basins. Moreover, sixteen wind

Many difficulties associated with modern land man-

energy projects have been constructed in the state since

agement are attributable to rising demands on limited

1999, primarily south of Casper and along the Inter-

resources. This is occurring at a time when climate

state 80 corridor (fig. 18.8). More are envisioned. The

is becoming less favorable than when many societal

Chokecherry and Sierra Madre Wind Energy Project,

expectations and cultural traditions were developed.

now on the planning boards, anticipates 1,000 turbines

Ecosystem management is an attempt to achieve sus-

spread over an area of about 350 square miles of sage-

tainable land management by integrating existing eco

brush-dominated shrubland near Rawlins. Aside from

logical information with the very real and difficult

fragmentation concerns, when vertical structures are

challenges faced by managers. Effective eco


system

involved, such as the towers required for transmission

management also transcends artificial boundaries.

lines, prairie chickens and sage-grouse instinctively shy

Much of the best wildlife habitat, for example, occurs

away from the area to avoid the risk of predation by rap-

on private lands, whose owners depend on public land

Using Western Landscapes

Fig. 18.8. Wind energy projects also contribute to the fragmentation of habitat and open space. Towers for the turbines and
associated power lines provide perches for raptors that prey on
birds and small mammals more often than they would without

the perches, thereby extending the effects of this kind of


industrial development. Power companies strive to locate power
lines along corridors to minimize their visual and ecological
impacts. Photo Dave Showalter, Lighthawk Aerial Support.

resources (such as ranchers who have grazing leases, or

trary, as animals, plants, water, nutrients, sediments,

landowners with property that is highly valued because

energy, and pollutants move or disperse across them

it is adjacent to public land).

(see chapter 1). The size of the area needed to imple-

The ecosystem management concept reinforces

ment an ecosystem management approach depends on

such commonly accepted practices as environmental

what ecological issues the management plan is designed

impact assessment, public participation in the develop-

to address. For example, the area included in the Greater

ment of management plans, and monitoring of resource

Yellowstone Ecosystem was determined to a large extent

values to ensure that management practices are work-

by considering the land used by the widely dispersed

ing as intended. Goals include managing for long-term

elk and grizzly bear populations (see chapter 15). In all

sustainability, which means maintaining productive

cases, various landowners are involved; collaboration is

capacity of the soil (as mandated in the Multiple Use

needed.

Sustained Yield Act) and preventing species extinction

Understanding that some landowners might feel

wherever possible (as mandated by the Endangered

threatened by this approach, proponents of ecosystem

Species Act). The nature of ecosystems in the region

management strive to achieve sustainability while avoid-

certainly changed dramatically during the Holocene

ing adverse impacts on people who depend on the land-

and before (see chapter 2). Now, however, with greater

scapes of the region and who, indeed, can be important

societal demands on natural resources, combined with

stewards of the land. Enabling them to prosper reduces

habitat fragmentation and accelerated climate change,

the likelihood of land sales that could lead to sub

precautions seem prudent. Ecosystem management is a

divisions or other uses that are not helpful. Indeed, mak-

step toward being more cautious.

ing a living in a region with a semi-arid, cool climate

How much land is necessary for effective ecosystem

can be difficult. The most appropriate livelihoods and

management? Human-defined ecosystem boundaries,

management practices may not have been achieved dur-

like political boundaries, are always somewhat arbi-

ing a mere century of trial and error. Though difficult

313

314 Sustainable Land Management


Fig. 18.9. Monitoring water
quality provides an indication
of how successfully ecosystem
services on the watershed are
being maintained. Photo of
the Sweetwater River by Tony
Ferlisi.

to contemplate, some changes could lead to an economy

by native plants and that provide various amenities

that is more stable than the region has experienced since

and much of the Wests characterare now being

the late 1800s. Ecosystem management is an approach

fragmented.14 Some residents are convinced that such

that motivates willing participants to explore their

trends are not good for the Wests economy. For exam-

options more broadly than may have been done before.

ple, Montana State University ecologist Andrew Hansen


and his associates wrote, local communities may best
be able to promote economic growth by maintaining

Amenities and Ecosystem Services


A dollar value easily can be placed on water, meat,
wood, oil, gas, coal, and other commodities on which
the livelihoods of many residents depend. In the New
West, however, both old-timers and newcomers are
placing considerable value on open space, clean air,
free-flowing streams with clear water, and abundant
wildlifeincluding nongame species (figs. 18.9 and
18.10). For many years there seemed to be little conflict. Roads and ranches here and there contributed to
the pastoral scene that many enjoy; industrial developments and rural subdivisions were viewed as progress.

the natural amenities that are so attractive to new residents and businesses. If this is true, policies that favor
timber harvesting over scenery, mining over water quality, or intensive livestock grazing over wildlife habitat
might actually inhibit rather than expand economic
growth.15
Ecosystem services are also becoming more highly
valued, as discussed in previous chapters. The costs of
replacing them with new technology, where that is possible, can be high. Herb Bormann, long-time ecologist
at Yale University, wrote in 1976 that, to the degree eco
system services are lost,

Now such developments are becoming progressively

they must be replaced. . . . We must find replace-

closer to each other. Opportunities to appreciate open

ments for wood products, build erosion control

space decline with too many wind energy projects, oil

works, enlarge reservoirs, upgrade air pollution con-

and gas wells, pumping stations, ranchettes, power

trol technology, install flood control works, improve

plants, and power lines extending to the horizon.

water purification plants, increase air conditioning,

Widespread protective landscapesthose dominated

and provide new recreational facilities. . . . These

Using Western Landscapes


Fig. 18.10. Pronghorn are
now abundant in Wyoming,
numbering about 500,000,
but they were threatened
with extinction in the early
1900s from overhunting. The
conservation of this species
and some others has been a
success. Photo by Ken Driese.

substitutes represent an enormous tax burden, a

funding for such payments, the conservation banks

drain on the worlds supply of natural resources, and

sell conservation credits, usually to corporations seek-

increased stress on the natural systems that remain.

ing ways to mitigate the potentially adverse impacts of

. . . Increased consumption of fossil energy means

other developments they sponsor. The motivation for

increased stress on natural systems, which in turn

mitigation may be driven by corporate philosophy

means still more consumption of fossil energy to

which appeals to many consumers and investorsor by

replace lost natural functions if the quality of life is

regulatory mechanisms.

to be maintained.16
Bormann recognized, more than 35 years ago, the
economic value of using solar-powered ecosystems to
provide numerous services, including erosion control,
clean water, recreational opportunities, and habitat for
wildlife. Today we can add other services, such as carbon
sequestration and resistance to invasive plants. Fossil
fuels have put Wyoming in the position of subsidizing
new developments through the Wyoming Permanent
Mineral Trust Fund, made possible by royalties and severance taxes on oil, natural gas, and coal. However, such
nonrenewable resources can be used only as the basis

The federal government has developed similar programs. A good example is the Conservation Reserve
Program, whereby farmers are compensated for planting perennial plants on erodible soils that had been cultivated for annual crops. Erosion is reduced, there is less
dust in the air, water draining from such lands is cleaner,
and wildlife habitat is improved. Similar programs also
provide economic incentives to discourage the draining
of wetlands and the plowing of still-intact grasslands.
In this way, consumers from across the nation pay for
highly valued ecosystem services provided by participating landowners.

for a transitional economy. Recognizing environmental


constraints is fundamental to achieving sustainability.
The challenge is to find ways by which people can

Development by Design and Climate Change

profit from selling ecosystem services. Progress is being

The perceived antagonism between economic develop-

made. Numerous corporations across the nation have

ment and conservation has led to various approaches for

created conservation banks that are now federally regu-

setting priorities that are complementary to ecosystem

lated and employ experts in conservation finance.17 If

management. One example, known as Development by

landowners commit to making significant ecological

Design, was first articulated by The Nature Conservancy

improvements to their land, and if an objective third

(a nongovernmental organization) and stems from the

party confirms such improvements, payments are made

environmental impact statement process mandated by

to the landowner, entirely from private funds. To obtain

the National Environmental Policy Act of 1969.18 This

315

316 Sustainable Land Management

approach, now used by various entities, calls for assem-

previously degraded site, the purchase of conservation

bling the best available biological and ecological infor-

credits from a conservation bank, or the purchase of

mation about sites where a development is proposed.

conservation easements that prevent habitat loss on

That information is then displayed on maps using Geo-

other sites that probably would be the location of future

graphic Information Systems. The severity of anticipated

developments and the associated consequences. The

environmental impacts are assessed in a similar man-

approach is not perfect; population sizes of some sensi-

ner. After careful analysis, the impacts and conflicts are

tive species will decline. However, the process is cur-

identified.

rently viewed as minimizing adverse effects and may

But Development by Design does not stop there.

prove to be generally acceptable to diverse stakeholders.

The next step is to formulate a strategy for avoiding,

The goal is to stop declines in biological diversity. Com-

minimizing, or mitigating the adverse effects. Actions

pensation from a conservation bank is available for

are then taken, and the result is monitored to measure

landowners who can assist with the mitigation pro-

whether the plan was successful. Based on the results,

cess, as described previously. Also, for landowners who

plans developed in the future can be improveda

collaborate with U.S. Fish and Wildlife biologists in

decisionmaking process known as adaptive manage-

developing a conservation plan for sensitive species on

ment. The best scientific information is used, and

their property, there is the potential for a safe harbor

the analyses are peer reviewed, as required with any

agreement.21

science-driven approach.19 The process includes willing participants. The first application in Wyoming

Over the years, land managers have learned how to

was on the Jonah gas field south of Pinedale. To start

restore degraded ecosystems by fostering the growth

with, several corporations contributed some $25 mil-

of desired plant species, closing some roads, reduc-

lion to a fund administered by the Jonah Interagency

ing recreational impacts, applying new approaches

Office, which is overseen by a committee composed of

for conserving threatened species, and a variety of

specialists with the Wyoming Game and Fish Depart-

other initiatives. Ecosystem services often have been

ment, Wyoming Department of Agriculture, Wyoming

restored. Success is always more likely when the climate

Department of Environmental Quality, and the Bureau

is favorable and familiar. Now, however, the climate is

of Land Management. Working in collaboration with

changing in ways that are affecting land management

the Wyoming Wildlife and Natural Resources Trust,

practices. Also, through the widespread application of

established by the Wyoming State Legislature, the

fertilizers and nitrogen-rich effluents into air and water,

fund is used for various kinds of mitigation, including

ecosystems are becoming enriched with nutrients. The

habitat improvement.

environment of the future is likely to be more fertile as

According to Development by Design protocol, some

well as warmer and drier.

sites will have very high conservation value and the

Native plants and animals have evolved tolerances to

best strategy is to avoid the impact. When this happens,

the slowly changing climatic conditions of the past tens

developers are asked whether they could find another

of thousands of years. Can they evolve rapidly enough

place or method for their investment. If that is clearly

to thrive under the changes expected during the next

not possible, a plan is agreed upon to minimize the on-

century? Will mitigation or restoration be feasible with

site impact and mitigate the adverse effects. Mitigation

combinations of precipitation, temperature, and sea-

involves the use of offsets, that is, the reclamation or res-

sonal variations that have most likely never existed

toration of nearby habitat agreed to be in relatively poor

before? Will exotic species be favored over the natives?

condition. Ideally, by minimizing adverse on-site effects

Such topics have been discussed throughout this book.

and providing for offsets, there is a net gain in habitat for

Warmer temperatures and longer summers are likely

the sensitive species in the area. Mitigation is most likely

to extend the inevitable droughts. Fires already have

to succeed if it is done near the affected site.20

become more frequent; water for municipalities, agricul-

The extent to which mitigation is possible is con-

ture, and industry has become more expensive, and the

troversial, but it might involve active reclamation of a

reallocation of limited water supplies is controversial.

Using Western Landscapes

The carrying capacity for some rangelands will decline,

for the past few centuries or more are likely to be sorted

and the cost of hay is likely to go up. Large herbivores,

into new ecosystems. 23 Locally, will such species adapt

such as elk and deer, may have difficulty surviving long

rapidly enough to the new climate? Will they become

winters if forage availability on their summer ranges is

rare or extinct because of insufficient habitat regionally?

insufficient due to drought, or if extraordinary stresses

Will traditional forms of land use become impractical?

during migration reduce their limited fat reserves.22

A prudent strategy is to conserve the full range of bio-

Native ecosystems that have attracted and helped sus-

logical diversity wherever possible. If that valuable but

tain westerners for nearly 200 years will change, some-

vulnerable resource can be saved, and water resources

times in ways that are uncomfortable.

can be used more efficiently, then future generations

Some earth scientists have projected that, with climate change, species that lived in the same ecosystems

are more likely to continue benefiting from the amenities and services provided by western ecosystems.

317

This page intentionally left blank

Epilogue

The changes that have taken place in western North

development tended to be downplayed; the prevailing

America over tens of millions of years are staggering to

disposition has been simply to use resources. To illus-

comprehendthe sudden extinction of dinosaurs,

trate, an older building on the University of Wyoming

the slow evolution of bison and pronghorn from now-

campus is engraved with the words Strive onthe

extinct ancestors, the gradual transition from humid

control of nature is won, not given. Another building,

tropical forests to sagebrush, and the expansion and

completed 24 years later, in 1950, greets students with

retreat of glaciers. Substantive changes also have taken

The foundation of agriculture is not rooted in soil but

place during the past 15,000 years, since the ancestors

rather in the vision and attainment of men.

of Native Americans arrived. And now, modern civiliza-

But adverse environmental impacts are now so per-

tion has accelerated the rate of change in ways that we,

vasive that they can no longer be ignored. The experi-

the authors of this book, find troubling.

ence of land managers and the research of scientists are

The mountains and plains of the West make up an

providing insights on how to proceed, and Congress has

area for which many residents and nonresidents feel

passed legislation to help ensure the nationwide appli-

responsible. Considering that roughly half of the land

cation of certain restrictions. These laws complicate the

in the region belongs to the American public at large,

lives of resource managers and landowners. Thats inev-

decisions are not made solely at the local level. More-

itable as human pressures increase and many resources

over, the various groups involved, often referred to as

become less readily available. Resource management

stakeholders, have differing perspectives and values.

and economic development will never be as simple as

Debates commonly focus on land management prac-

they seemed to be 50 or 100 years ago. Classroom dis-

tices, the effects of unavoidable disturbances, and the

cussions now probe this complexity, providing a more

reasons that ecosystems respond the way they do. Fre-

realistic perspective than conveyed by a few words

quently the discussions consider strategies for benefit-

etched in stone on buildings constructed when the

ing well into the future from the amenities, services,

worlds population was less than half what it is today.

and natural resources that native ecosystems provide.

Providing a high quality of life for so manyone that

Modern science, technology, and our economic system have greatly augmented the availability of many

can be sustained for the benefit of future generations


is a formidable challenge.

natural resources, often improving the enjoyment of

Wyoming and the Rocky Mountain region in general

life in ways that once seemed sustainable for many

must contribute to the fossil fuels, minerals, wood fiber,

generations. Until recently, the limits to growth and

food, and other resources required for the great human

319

320 Epilogue

endeavor now under way. While providing these com-

of the plants and animals existed a million years or more

modities, however, we should not lose sight of the non-

before Homo sapiens arrivedand we are still learning

material qualities in which this region is so richopen

how to make a living from rugged western landscapes. As

spaces, abundant wildlife, free-flowing rivers, and others.

Aldo Leopold wrote in 1938, the oldest task in human

It is critical that the land be used with care and vision,

history [is] to live on a piece of land without spoiling it.1

so that less tangible but equally valuable resources and

Learning to live gently and sustainably, to be good stew-

services are not lost needlessly. Humans have been a pres-

ards, requires an understanding of both human nature

ence in this part of the biosphere for a short timemost

and the nature of ecosystems.

Appendix A Conversion Table

To convert

Multiply by

LENGTH
Inches to centimeters

2.54

Feet to meters

0.304

Miles to kilometers

1.609

Millimeters to inches

0.0394

AREA
Square feet to square meters

0.0929

Acres to hectares

0.405

Square miles to square kilometers

2.590

MASS
Ounces to grams

28.4

Pounds to kilograms

0.454

Tons (2,000 lb) to metric tons

0.907

VOLUME
Cubic feet to cubic meters
Cubic feet to liters

0.0283
28.3

Quarts to liters

0.946

Gallons to liters

3.78

TEMPERATURE
Fahrenheit to Celsius

5/9 (F 32)

Celsius to Fahrenheit

9/5 (C + 32)

(A difference of 1F is 0.556C; a difference of 1C is 1.8F)


ENERGY
Calories to joules

4.19

321

Appendix B Characteristic Soil Types

Table B.1. Soil orders and subgroups characteristic of vegetation types in Wyoming and adjacent parts of
neighboringstatesa
Soil order
Location/vegetation type

Entisols

Inceptisols

Aridisols

Mollisols

Alfisols

UsHAI
UsNAI

ArAUO

Black sagebrush

TyCOI
TyHAI

Big sagebrush

TyHAI
BoHAI

TyTFE
TyTPE

LOWLANDS
Mixed-grass prairie
Sagebrush steppe

Silver sagebrush
Riparian vegetation
Lowland

TyHAO

Foothill

CuHCO

Montane

HiCAP

Alpine

HiCAP

Saltbush

TyHAI
TyNAI

Mixed

TyNAI

Greasewood

BoNAI
TySOI
TyGOI

Basin grassland

UsHAI
UsNAI

Desert shrublands

322

Characteristic Soil Types 323

Soil order
Location/vegetation type

Entisols

Inceptisols

Aridisols

Mollisols

Alfisols

Sand dunes (stabilized)

TyTPE

Badlands

TyTOE

TySOI

TyHDO

Playas
FOOTHILLS
Shrublands
Mixed
Mountain-mahogany

LiTOE

Threetip sagebrush

TyCOE

TyCCP

Mountain big sagebrush

TyHCO

Foothill grassland

ArHUO
ArAUO

Oak woodland

TyUOE

Juniper woodland

TyTOE

Ponderosa pine

LiTOE
LiUOE

TyHUA
TyHDA

Woodlands

Douglas-fir
Aspen

TyHAO

TyHDO

MOUNTAINS
Forests
Ponderosa pine

TyHDA

Lodgepole pine

TyCCP

TyHCA

Engelmann sprucesubalpine fir

TyCCP

TyHCA

Douglas-fir

MoHCA
TyHDA

Aspen

TyHCO

AqHCA

Dry

TyHCO

Moist

AqHCO

Wet

HiCAP

TyCAO

Fellfield

HuDCP

Turf

HuDCP
TyDCP

Wet

TyCAE

Subalpine meadows

Alpine tundra

Source: Courtesy of Larry Munn.


a

Subgroup abbreviations are defined in table B.2. A dash indicates that the soil type is absent or not so common. For descriptions of soil

orders and subgroups, see http://soils.usda.gov/technical/classification/tax_keys/.

324

Characteristic Soil Types

Table B.2. Soil subgroup name abbreviations


AqHCA

Aquic Haplocryalfs

TyCAO

Typic Cryaquolls

TyHUA

Typic Haplustalfs

AqHCO

Aquic Haplocryolls

TyCOE

Typic Cryorthents

TyNAI

Typic Natrargids

ArAUO

Aridic Argiustolls

TyCOI

Typic Camborthids

TySOI

Typic Salorthids

ArHUO

Aridic Haplustolls

TyDCP

Typic Dystrocryepts

TyTFE

Typic Torrifluvents

CuHCO

Cumulic Haplocryolls

TyGOI

Typic Gypsorthids

TyTOE

Typic Torriorthents

HiCAP

Histic Cryaquepts

TyHAI

Typic Haplargids

TyTPE

Typic Torripsamments

HuDCP

Humic Dystrocryepts

TyHAO

Typic Haplaquolls

TyUOE

Typic Ustorthents

LiTOE

Lithic Torriorthents

TyHCA

Typic Haplocryalfs

UsHAI

Ustic Haplargids

UsNAI

Ustic Natrargids

LiUOE

Lithic Ustorthents

TyHCO

Typic Haplocryolls

MoHCA

Mollic Haplocryalfs

TyHDA

Typic Hapludalfs

TyCAE

Typic Cryaquents

TyHDO

Typic Hapludolls

Source: Courtesy of Larry Munn.

Notes

Chapter 1: Introduction
1. Hay, including alfalfa, is by far the most abundant crop in
Wyoming and is grown on about two-thirds of the cultivated
land. Also common are sugar beets, corn, beans, wheat, and
barley. Livestock grazing is widespread. Timber harvesting in
mountain forests is a much smaller industry than agriculture.
About 29 percent of the gross state product in 2011 was generated from natural gas, oil, coal, trona, and bentonite, compared
to about 6 and 3 percent, respectively, from tourism and agriculture. See Hamerlinck et al. 2013.
2. Elk are numerous in many mountain ranges, and a small herd
persists in the desert northeast of Rock Springs, near Steam-

6. For further discussions of ecosystem services and their marketable values, see Costanza et al. (1997), Daily and Ellison (2002),
Clark (2007), Burke et al. (2008), Perrings et al. (2011), and
Hansen et al. (2013). For the effects of climate change on biodiversity and ecosystem services, see Frontiers in Ecology and the
Environment (vol. 11, issue 9).
Chapter 2: Landscape History
1. Snoke (1993) provides a detailed synthesis of the geologic history of Wyoming; Johnson and Raynolds (2003) provide an
illustrated overview; for a detailed geology map, see Love and
Christensen (1985).

boat Mountain and the Killpecker Sand Dunes. Grizzly bears

2. Tidwell 1975; some fossils were subsequently lost by erosion.

and wolves now occupy much of their former habitat in north

3. Most Cretaceous volcanism occurred outside the area that

western Wyoming.
3. The change that occurs after ecosystem disturbances is known

would become Wyoming.


4. Lillegraven and Ostresh 1988; Snoke 1993.

as secondary succession. Communities and ecosystems are con-

5. B. Breithaupt, pers. comm.

stantly changing in one manner or another.

6. In 2006, south of Laramie, the fossils of five dinosaurs were dis-

4. Landscapes may be large or small but are spatially patchy. Ecol-

covered in a trench dug for a natural gas pipeline. Three genera

ogists interested in grasshoppers focus on smaller landscapes

were found (Allosaurus, Apatosaurus, and Comarasaurus) in the

than those studying bears and humans. See Turner et al. (2001)

Morrison Formation of the Jurassic (K. Trujillo, pers. comm.).

for an overview of landscape ecology, and Chapin et al. (2012)

The ancestors of pronghorn appeared in the fossil record of

for an overview of ecosystem ecology.

the Miocene, about 20 million years ago; bison did not appear

5. Various organizations now exist for studying and monitoring the

until late in the Pleistocene, less than 300,000 years ago. Bison,

status of rare and endangered species. They employ conservation

mammoths, mastodons, camels, and horses were numerous by

biologists and are found in every state. The Wyoming Natural

20,000 years ago; see Dott and Batten 1981; Axelrod 1985; Hart

Diversity Database, affiliated with the University of Wyoming

2008.

(http://www.uwyo.edu/wyndd/about/index.html), provides ser-

7. Tidwell 1975.

vices to industry, municipalities, conservation organizations,

8. Formerly known as the Cretaceous-Tertiary (K-T) or Cretaceous-

and federal and state agencies. The U.S. Fish and Wildlife Ser-

Paleogene (K-Pg) Extinction Event.

vice administers the Endangered Species Act and provides lists

9. Schulte et al. 2010; one crater has been found in the present-

of endangered, threatened, and candidate species by state. The

day Gulf of Mexico, just north of the Yucatan Peninsula. Other

Wyoming website is at http://www.fws.gov/wyominges/Pages/

explanations include extremely large volcanic eruptions at

Species/Species_Endangered.html.

about the same time in India.

325

326

Notes to Pages 1323


10. Robertson et al. 2004.

26. Petit et al. 1999.

11. Gingerich and Clyde (2001) and Clyde (2001) list more than

27. Mears 1997.

150 genera of mammals known from Paleocene-Eocene fossils

28. Tidwell 1975.

in the Bighorn Basin; Lofgren et al. (2004) list the genera of

29. Current estimates range from 16,000 to 13,500 years ago: http://

Paleocene mammals for North America, based on many fossils

www.nytimes.com/2012/05/20/opinion/sunday/who-arrived

collected in Wyoming and nearby.

-in-the-americas-first.html. Gardner (2008a) provides a sum-

12. Tidwell 1975; Riedl (1959) describes the fossils of the Shirley

mary of human history for the region, focusing on the Red

Basin; Dorf (1964) reports on the fossil trees in Yellowstone

Desert, and Kornfeld et al. (2010) provide an overview for the

National Park; Wing et al. (2005) describe rapid global warming

Great Plains and Rockies; see Janetski (1987) for human history

and vegetation change in the Bighorn Basin during the PETM;

in northwest Wyoming.

Kunzig (2011) includes an artistic rendition of what the plants

30. The Colby Mammoth Site is located in the Bighorn Basin,

and animals in the Bighorn Basin looked like during this time,

near Greybull; see http://www.wyohistory.org/encyclopedia/

with estimates of temperature changes; Eberle and Greenwood


(2012) review the Eocene flora of the Arctic.
13. Pocknall and Flores 1987; the Wyoming State Geological Survey

colby-mammoth-site.
31. Reher and Frison 1980; see http://www.aaanativearts.com/
tribes-by-states/wyoming_tribes.htm.

website. The coal of the Powder River Basin is in the Fort Union

32. Working from boats, paleoecologists take vertical cores from

and Wasatch formations, deposited during the Paleocene and

the bottom sediments and are able to identify the plants that

Eocene epochs, respectively. Coal is found in other parts of


Wyoming as well, but often is under deep sediments or exists
only in thin seams.
14. Wing 1981; Gemmill and Johnson (1997) report on fossils
found in the Great Divide Basin that date back about 58 million

produced the preserved pollen.


33. Baker 1986; Walker 1987; Whitlock 1993; Shuman et al. 2009;
Mumma et al. 2012; Krause and Whitlock 2013.
34. Baker 1986; Whitlock 1993; Licciardi and Pierce 2008; Shuman
et al. 2009; Mumma et al. 2012; Krause and Whitlock 2013.

years (the late Paleocene); Barnosky (1984) described similar

35. Walker 1982, 1987.

vegetation patterns for the Jackson Hole area in the Miocene.

36. Reider 1983, 1990; Reider et al. (1988) describe grassland soils

15. Zachos et al. 2008; McInerney and Wing 2011; though a global

where forests now grow in the mountains; Marcott et al. (2013)

phenomenon, much of the data for the PETM was obtained in

conclude that the present-day global temperature is lower than

the Bighorn Basin of Wyoming, as reviewed by Kunzig (2011).

during the early Holocene but is likely to be higher by the year

16. McInerney and Wing 2011.


17. Jansen et al. 2007.

2100 (see chapter 3).


37. Minckley et al. (2012) present evidence for an earlier warm

18. A portion of the initial Medicine Bow Mountains was later sepa-

and dry period, based on hydrologic fluctuations, archeol-

rated by faulting and is known today as the Sierra Madre; see

ogy, and sand dune activity. See also Forman and Pierson

fig.1.2.

2003; Mayer and Mahan 2004 (Killpecker Dunes); Forman et

19. Sedimentary strata formed at the bottom of these lakes are now
in the Green River Formation, which is surrounded and overtopped by the lake-bordering fluvial and swampy deposits of
the Bridger Formation; Grande 1998; Thompson 2008.

al. 2005; Hart 2008; Halfen et al. 2010 (Casper Dune Field);
Mason et al. 2011.
38. Jackson et al. 2005; see M. E. Buskirk (2009) for a description of
packrat behavior and their nests.

20. Thompson 2008 (p. 111).

39. Lyford et al. 2003.

21. Love et al. 2003.

40. Norris 2006; Lesser and Jackson 2012.

22. Axelrod 1985; Leopold and Denton 1987; Mears 1993.

41. Gray et al. 2006.

23. Two periods of extensive glaciation are widely recognized in

42. Lyford et al. 2002; Jackson et al. 2002.

Wyoming: the Pinedale (about 30,000 to 14,000 years ago)


and the Bull Lake (about 150,000 to 100,000 years ago). For

43. Drying in the early Holocene also is indicated by the pollen


record, as noted previously.

the history of glaciation, see Mears (1997, 2001), Pierce (2003),

44. Fall et al. 1995.

and Licciardi and Pierce (2008). The Pinedale glacial period is

45. Pine expansion is coincident with active sand dunes and low

known in midwestern states as the Wisconsin glacial period.


24. Grasso (1990) describes Lake Wamsutter, located mostly north

water levels in lakes and riverstwo indicators of a dry or drying climate (B. Shuman, pers. comm.).

of Wamsutter, Wyoming. At times it was more than 180 feet

46. Jansen et al. 2007; Gardner 2008b; Marcott et al. 2013.

deep and covered 700 square miles, including Separation Flat,

47. Gray et al. 2004, 2006.

Chain Lake Flat, and Battle Springs Flat. About 10,000 years ago

48. Reider et al. (1988) conclude from soil profile data that fluctua-

the Killpecker Sand Dunes developed on top of a portion of the


lakebed.

tions occurred from about 3,000 to 200 years ago.


49. Some scholars assume that Native Americans had almost no

25. Snoke (1993) and Good and Pierce (2002) describe the forma-

impact on western North American ecosystems, because their

tion of the Yellowstone Plateau, caused largely by volcanic erup-

populations were low and their technology was limited. Others,

tions at about 2.0, 1.3, and 0.6 million years ago.

such as Charles Mann (2005), argue that Native Americans were

Notes to Pages 2337


numerous and had significant effects. Mann concluded that
European diseases decimated Indian populations, which led to
the erroneous impression that their numbers were low. Thomas
Vale (2002) concluded that the greatest human impacts prior
to arrival of Europeans could have been near popular camping
locations on the plains. See also Kornfeld et al. 2010.
50. Gardner 2008b; Beauvais (2008b) indicates that horses were in
Wyoming in the 1600s; Proulx (2008) reviews the history of
EuroAmericans in the Red Desert; Janetski (1987) writes about
Native Americans and the horse in Yellowstone.
51. Proulx 2008.
52. The first authenticated EuroAmerican in Wyoming arrived in
the Powder River Basin in 1805Franois Antoine Larocque,
a French-Canadian fur trader. The La Vrendrye brothers may
have been in the northeast part of the state in 17421743, but
that is not documented.
53. Hart (2008) reviews the history of land use for the western
Great Plains; see also Blair 1987; Crowe 1990; Gardner 2008a.
54. Gardner 2008a; Persico and Meyer (2009) conclude that the significance of valley filling by beaver and gully formation after
dam failure may be exaggerated in some areas. Earlier journals
suggest that gulleys were common before beaver pelts became
highly prized.
55. Reher and Frison 1980; Vorebuffalojump.org.
56. Occasional bison herds of tens of thousands were seen as late as
the 1870s, but by 1882 many Native Americans were starving,
because the bison were so rare; by 1895 only about 800 bison

Chapter 3: Present-Day Environments and Climate Change


1. Curtis and Grimes 2004; the climate of the shortgrass prairie is
described by Pielke and Doesken (2008); the highest precipitation tends to occur at mid-elevations in the mountains.
2. Smith and Geller 1979.
3. Mack and Thompson 1982.
4. Webb et al. 1978; Joyce 1981; plant growth is also known as
net primary productivity and is commonly expressed as grams/
square meter/year.
5. Moist air is brought to eastern Wyoming when the atmospheric
pressure to the south is low. Air flows are counterclockwise at
such times, which brings moist air from the Gulf coast north
and west to Wyoming.
6. Shinker et al. 2010; Wise 2010.
7. Martner (1986) presents data that show how spring and early
summer precipitation is more characteristic of the eastern
plains, whereas the western two-thirds of the state often has
a more even distribution of precipitation during the year (and
a larger proportion that occurs in the winter as snow). See also
Driese et al. 1997; Bates et al. 2006; Schlaepfer et al. 2011.
8. Depending on how drought is defined. Martner (1986) identified droughts in the mid-1930s, mid-1950s, and the early 1960s.
9. Curtis and Grimes 2004 (p. 97); Gray and Andersen 2009;
Woodhouse et al. (2006) observed that Colorado River water
allocation compacts were approved during one of the wettest
periods in the past 500 years; see also Wise 2012.

remained in North America. The last wild bison in Wyoming,

10. This quantity is known as the adiabatic lapse rate; Curtis and

outside of Yellowstone National Park, died in the Red Desert in

Grimes (2004) estimated a drop of 5.5F per 1,000 feet for Wyo-

1889 (Beauvais 2008).


57. The bison of the western Great Plains once numbered in the
millions (Cushman and Jones 1988; Hart 2001).

mings mean maximum temperature in July.


11. Martner 1986.
12. Martner 1986; Ostresh et al. 1990.

58. Dorn 1986.

13. Martner 1986; Curtis and Grimes 2004.

59. Larson 1977; Schullery and Whittlesey 1992; Kornfeld et al.

14. http://www.windpoweringamerica.gov.

2010.
60. Larson 1977; the effects of grazing by domestic livestock along

15. Sala et al. 1988; semi-arid is defined as having an annual precipitation of less than 15 inches.

the Oregon Trail might have been minimal if the animals were

16. Ganskopp 1986.

moved westward during midsummer or fall and if the animals

17. Branson et al. 1967, 1970, 1976. Common salts in the region

grazed the rangeland for only a short time. See also Wentworth

include sodium and magnesium chloride; sodium, calcium, and

1948; Mitchell and Hart 1987; Miller et al. 1993; http://www

magnesium sulfate; and bicarbonates of calcium and magne-

.wyohistory.org/essays/wyoming-cattle-boom.

sium. Saline soils can be defined as any soil with concentra-

61. Mitchell and Hart 1987 (p. 5).

tions of soluble salts that produce an electrical conductivity

62. In January 2011 about 1.3 million cattle and 360,000 sheep

of 4 dS/m or more. Alkalinity refers to soils in which the less

grazed on Wyoming rangelands (USDA National Agricultural

soluble calcium carbonate is important (pH 7.88.2), and sodic-

Statistics ServicesWyoming). Sheep numbers peaked in 1908

ity implies that sodium occupies 15 percent or more of the soil

at about 6 million. See also Abbott and Smith 1955; Larson

adsorption sites. Plant survival on such soils is dependent on

1978.

an ability to maintain cells with a solute concentration higher

63. Johnson 1987.

than that of the soil solution. If this is not the case, water flows

64. No trees are seen in early landscape photos along Crow Creek

out of the plant into the soil, creating water stress even though

near Cheyenne.

the soil is moist.

65. Hart and Laycock (1996) provide a bibliography of repeat-

18. The grazing of large herbivores (such as bison, elk, cattle, and

photo studies in the Great Plains and central Rocky Mountains.

sheep) becomes a disturbance in some grasslands only if too

Meagher and Houston (1998) show repeat photos in Yellow-

many animals are concentrated for too long in an area that is

stone National Park. See also Amundson 1991.

too small.

327

328

Notes to Pages 3756


19. Reported in Thwaites 1905 (p. 311).
20. See Wagner 2009 and online reports of the Intergovernmental
Panel on Climate Change (http://www.ipcc.ch), the National

4. Patten 1968; Starr 1974; Phillips 1977; Chadde et al. 1988;


Baker 1989a,b; Kittel 1994; Girard et al. 1997; Walford et al.
2001.

Climate Assessment program (http://www.globalchange.gov),

5. Houston 1967; Miller 1979.

and The Nature Conservancy (http://www.climatewizard.org/).

6. Akashi 1988; Friedman et al. 1996; Scott et al. 1997; Cooper et

21. See previous note.


22. Shuman 2012.
23. See Shuman 2012; chapter 12.
24. Wagner 2009; Pederson et al. 2010.

al. 1999.
7. Everitt 1968; Irvine and West 1979; Akashi 1988; Cooper et al.
1999; Hultine et al. 2010a,b; Rood et al. 2010.
8. Allred 1941; American elm occurs primarily in the Black

25. Naftz et al. (2002) examined oxygen isotope data in ice cores

Hills and has been killed in most areas on the Great Plains

taken from glaciers in the Wind River Range and concluded

by the introduced Dutch elm disease. Lanceleaf cotton-

that air temperatures had increased about 6F from the mid-

wood is a hybrid between plains cottonwood and narrowleaf

1960s to the early 1990s (about 9F from the mid-1800s to the


early 1990s). See also Naftz et al. 2009.
26. Shuman 2012 (p. 443).

cottonwood.
9. Raynolds 1868 (p. 83); Sacket 1877 (pp. 67).
10. Sun (1986) and K. Johnson (1987) write about the Sweetwater

27. Since Shumans study was completed, there has been one year

River; Cooke (1857, p. 402) about Crow Creek. Dorn (1986)

with high snowfall and one with low snowfall; the trend con-

reviews the journals of early explorers in Wyoming. Widely

tinues (B. Shuman, pers. comm.).

scattered groups of narrowleaf cottonwood were noted along

28. Woodhouse et al. 2006; Lundquist et al. 2009; Mote 2009;


Pederson et al. 2010, 2011; Shinker et al. 2010; Leppi et al. 2011;
Rice et al. 2012.

the Laramie River, at higher elevations in the southern part of


the Laramie Basin.
11. Based on repeat photos in K. Johnson (1987), Baker (1990),

29. Laramie Boomerang, October 19, 2012; more than half (57 per-

Knopf and Scott (1990), Friedman et al. (1996), and others;

cent) of Wyoming had extreme or exceptional drought, and 98

the distribution of cottonwood depends on flooding in some

percent had at least moderate drought.


30. Long-term precipitation data are available from the National

places, but floods also can damage young trees.


12. Akashi 1988.

Climatic Data Center (http://www.ncdc.noaa.gov/oa/ncdc

13. Sedgwick and Knopf (1989) conclude that cottonwood seedling

.html); downward trends are suggested in the Green, Upper

mortality along the South Platte River was caused by insufficient

Platte, Wind, and Yellowstone river basins, but not else-

water due to depth of the water table and by competition from

where in Wyoming. See globalchange.gov for projections of

nearby herbaceous plants. They suggest that cottonwood could

the National Climate Assessment and Development Advisory

be replaced by ash, Russian olive, or eastern juniper. In contrast,

Committee.
31. Anderegg et al. 2010; Stocker 2013. Carbon dioxide concentra-

along the North Platte River in Nebraska, cottonwood regeneration is now abundant and apparently does not depend on

tion in Earths atmosphere is now at or near 400 ppm (Science

the formation of point bars. Woodlands are expanding (W. C.

News, Dec. 28, 2013); it was 280 ppm in pre-industrial times.

Johnson, pers. comm.), reducing the amount of habitat avail-

32. Naftz et al. 2002; Woodhouse et al. 2006; Lundquist et al. 2009;

able for the thousands of sandhill cranes migrating through

Mote 2009; Pederson et al. 2010, 2011; Leppi et al. 2011; Rice et

central Nebraska in late winter and early spring. See also Fried-

al. 2012.

man et al. 1996.

33. Gray and Andersen 2009 (p. 12).

14. Weynand et al. 1979; Miller 1979; Merritt and Cooper 2000;

34. Rehfeldt et al. 2006.

Miller et al. (1995) did a similar study of the Rawhide Wildlife

35. Finch 2012.

Management Area on the North Platte River; Coble and Kolb

36. Friggens et al. 2012 (p. 1).

(2013) studied the effects of regulated and unregulated flows on

37. Williams and Jackson 2007; Williams et al. 2007. For further

the Delores River in Colorado.

reading on the effects of climate change on biodiversity and

15. The importance of floods in riparian landscapes is the topic

ecosystem services, see volume 11 (issue 9) of Frontiers in Ecology

of special issues of BioScience (September 1998) and Ecological

and the Environment.

Applications (June 2001).


16. Houston 1967; narrowleaf cottonwood reproduction is rare or

Chapter 4: Riparian Landscapes


1. Olson and Gerhart 1982; Hubert 2004; Lohman 2004;
McK instry et al. 2004.

absent along the nearby Snake River below Jackson Lake dam,
probably because flood peaks are lower there.
17. Houston 1967; McBride and Strahan 1984; Youngblood et al.
1985; Cooper et al. 1999.

2. The Clean Water Act and Food Security Act place restrictions on

18. Boggs and Weaver (1994) observed this pattern in Montana,

the management of wetlands that are connected to navigable

with the trees sometimes dying after only 60 years; Akashi

waters, whether by surface or groundwater.

(1988) observed a similar pattern along the Bighorn River in

3. Blue spruce is common in the foothill riparian zone along the


Snake River, lower Hoback River, and upper Wind River.

north-central Wyoming; see Miller (1979) for a discussion of


woodlands on the eastern plains.

Notes to Pages 5766


19. Neff 1957.

Medicine Bow Mountains that had been used for floating rail-

20. Rutherford 1954; Call 1966; Skinner et al. 1984; Parker 1986;

road ties from 1868 to 1940. Altered stream channels are still

Hayes 2012.

apparent.

21. Bragg 2000.

41. Stohlgren et al. 1998.

22. Brown et al. 1996.

42. Crops are not considered to be invasive plants, because they

23. D. Randall (1983); McKinstry and Anderson (1999) found that

usually do not displace native plants without some kind of dis-

many land managers are now using beaver for the benefits they

turbance, such as plowing; weeds may or may not be invasive.

provide. See also Cooke and Zack (2008).

The English sparrow and Eurasian collared dove are invasive

24. Beaver are able to obtain construction material for their dams

birds. Except for cutthroat trout, all species of trout in Wyo-

and lodges over distances of 150300 feet (B. H. Smith, pers.

ming are introduced. Among invertebrates, the New Zealand

comm.); Call 1966.

mud snail and a large, stalk-forming diatom are invasive on

25. Skinner et al. 1984.

stream bottoms in northwestern Wyoming (R. Hall, pers.

26. Dorn 1986: Gardner 2008a.


27. Dorn 1986.
28. Haines 1965 (p. 51).
29. Platts 1981; Platts et al. 1983; Kauffman and Krueger 1984;
Schulz and Leininger 1990; Clary et al. 1992.
30. Sedgwick and Knopf 1991; as on the adjacent uplands, grazing
can be a minor factor in riparian landscapes if carefully managed and if it occurs late in the growing season.
31. Platts 1981, 1982; Kauffman et al. 1983; Platts et al. 1983; Kauff-

comm.).
43. Lesica and Miles 2001a,b, 2004a,b; Katz and Shafroth 2003;
Pearce and Smith 2007.
44. Simons and Seastedt 1999; DeCant 2008; Shah et al. 2010;
Mineau et al. 2011; Reynolds and Cooper 2011.
45. Carman and Brotherson 1982; Sher and Quigley 2013.
46. Sexton et al. 2006; Pearce and Smith 2007; DeWine and Cooper
2008, 2010; Merritt and Poff 2010; Nagler et al. 2011.
47. Young et al. 2004.

man and Krueger 1984; Stabler 1985; Brookshire et al. 2002;

48. Ladenburger et al. 2006; Glenn et al. 2012.

McKinstry et al. 2004; Anderson (2007) observes that reduc-

49. Shafroth 2005; Owens and Moore 2007; Nagler et al. 2011;

tions in willow cover near winter feedlots for elk caused a


decline in the willow flycatcher, MacGillivrays warbler, and fox
sparrow.
32. Patten 1968; Kay 1993; Singer et al. 1994; Beschta 2003; Wolf et
al. 2007; Beschta and Ripple 2010.
33. Sedgwick and Knopf (1991) found that livestock preferentially
grazed on recently fallen cottonwood leaves during late autumn
along the South Platte River in Colorado.
34. State Engineers Office, pers. comm.; the Jackson Lake dam on
the Snake River was finished in 1906, enlarging Jackson Lake
and creating the oldest large reservoir in Wyoming (Ostresh et
al. 1990).
35. Marston and Brosz 1990; 51 percent flows to the Missouri River,
32 percent to the Columbia River, 14 percent to the Colorado,
and 3 percent to the Great Basin (Ostresh et al. 1990).
36. A report on the Snake River Levee System is available at http://
www.jhalliance.org/Library/Reports/levee.pdf.

Hultine and Bush 2011.


50. Hultine et al. 2010a,b; the Mediterranean tamarisk beetle (Diorhabda elongata) and related species have been effective; Lesica
and Miles (2004b) found that 3 months of flooding kills salt
cedar in northeastern Montana. For water use by saltcedar compared to native species, see Sala et al. (1996), Smith et al. (1998),
and Sexton et al. (2006).
51. Lesica and Miles 2001a,b; DeWine and Cooper 2008, 2010;
Bunting et al. 2011.
52. Poff et al. 2011; invasive aquatic plants and animals have also
created ecological problems.
53. Allan and Flecker 1993; Banerjee et al. 2013.
54. Pederson et al. 2011; see chapter 3.
55. Shinker et al. (2010) identify periods when the North Platte
River was dry and discuss the implications for Wyoming water
rights agreements with Colorado and Nebraska.
56. Extensive mortality of cottonwood and other riparian trees was

37. Instream flow rights are also recognized on the Wind River

observed during the drought of the 1930s (Ellison and Wool-

Reservation; two rivers in Wyoming have been protected by the

folk 1937; Albertson and Weaver 1944; Friedman et al. 1996);

Wild and Scenic Rivers Act, namely, the headwaters of the Snake

narrow
leaf cottonwood may be more drought tolerant than

River and the Clarks Fork of the Yellowstone River. Less than

plains cottonwood.

1percent of Wyomings rivers and streams have been protected


by this Act and the states instream flow option combined. The
flows of designated instream flow segments are managed by the
Wyoming State Engineer. See Madson 2012; https://sites.google
.com/a/wyo.gov/seo/surface-water/instream-flow.

Chapter 5: Marshes, Playa Wetlands, Wet Meadows, and Fens


1. McKinstry et al. 2004; Mitsch and Gosselink 2007; Keddy 2010;
wetland values are acknowledged by the Clean Water Act and

38. Gribb and Brosz 1990.

provisions in farm bills passed by Congress (Ryan and Squillace

39. D. Brosz, pers. comm.; see chapters 5 and 17 for the importance

2004). See chapter 4 for a discussion of riparian wetlands. As

of irrigation to wetlands.
40. Wohl (2005) describes the history of human impacts on creeks
and rivers in the Rocky Mountains and the challenges of restoration; Young et al. (1994) identify 61 streams in or near the

defined by wetland ecologists, bogs and swamps are not found


in Wyoming.
2. U.S. Department of the Interior, Fish and Wildlife Service,
Washington, D.C. http://www.fws.gov/wetlands/.

329

330

Notes to Pages 6779


3. Copeland et al. 2009, 2010.
4. Peck and Lovvorn 2001; Lovvorn and Hart 2004; Sueltenfuss et
al. 2013.
5. A total of 434 plant species, subspecies, or varieties are considered by the Wyoming Natural Diversity Database as rare
enough in Wyoming to be vulnerable to extirpation (B. Heidel,
pers. comm.). Approximately 20 percent of these plants are
obligate wetland indicators; an additional 10 percent are facultative wetland indicators.
6. Lemly and Cooper (2011) study mosses; Hart and Lovvorn
(2000) report on the importance of algae.
7. Orabona et al. 2009; Lewis 2011.
8. Orabona et al. 2009.

24. Salt flats are a common feature in desert shrublands; see


chapter8.
25. Gage and Cooper 2012.
26. Gage and Cooper 2012.
27. Youngblood et al. 1985; Hansen 1995; Girard et al. 1997; see
chapters 4 and 13.
28. Peatlands are described by Rydin and Jeglum (2006), Mitsch
and Gosselink (2007), and Keddy (2010).
29. More than 200 species of bryophytes have been documented in
the Beartooth Mountains, mostly in fens, including Sphagnum
russowii and S. warnstorfii; see reports available at the Wyoming
Natural Diversity Database website (http://www.uwyo.edu/
wyndd/reports-and-publications/).

9. Twenty-nine species of birds are migrants, mostly shore-

30. Mattson 1984; Lemly 2007; Lemly and Cooper 2011; and reports

birds. Only eight species are considered year-round Wyoming

available at the Wyoming Natural Diversity Database website

residents.

(http://www.uwyo.edu/wyndd/reports-and-publications/),

10. Orabona et al. 2009; S. W. Buskirk, pers. comm.


11. Aquatic beetles, black flies, caddisflies, damselflies, dragonflies,

authored by B. Heidel, M. Jankovsky-Jones, G. Jones, S. Laursen,


E. Rodemaker, and R. Thurston.

fingernail clams, midges, mosquitoes, scuds, snails, water fleas,

31. Collins et al. 1984.

and water skippers; Wollheim and Lovvorn 1995; Gammonley

32. A minimum of 16 inches of peat is a common indicator of fens

2004.
12. For reviews of wetland adaptations, see Rydin and Jeglum
(2006), Mitsch and Gosselink (2007), and Keddy (2010).

in the United States.


33. Cooper and Andrus 1994.
34. See the 1992 report by W. Fertig and G. Jones and the 2000

13. The initial steps of respiration in all species in all environments

report by B. Heidel and S. Laursen, both available at the Wyo-

do not require oxygen, but a smaller portion of the energy

ming Natural Diversity Database website (http://www.uwyo

contained in food is available than in aerobic environments.


Mitsch and Gosselink (2007) and Keddy (2010) review the biogeochemistry of hydric soils.

.edu/wyndd/reports-and-publications/).
35. Lemly 2007; Lemly and Cooper 2011; fens on geothermally
altered sites on the Yellowstone Plateau, and in other parts of

14. Bernal and Mitsch 2012.

North America, are not easily classified using the nutrient-

15. Mitsch and Gosselink 2007.

availability gradient; such fens are acidic and similar to bogs,

16. Lewis 2011.


17. Gage and Cooper 2012.
18. Various studies have focused on wetlands at higher elevations.
They include Briggs and MacMahon (1983); Mattson (1984);

though nutrient concentrations tend to be higher than in bogs.


36. The mosses include Aulacomnium palustre, Drepanocladus aduncus, Philonotis montana, Plagiomnium cuspidatum, Ptychostomum
pseudotriquetrum, and several species of Sphagnum.

Youngblood et al. (1985); Chadde et al. (1988, 1998); Girard et

37. Lemly 2007; Lemly and Cooper 2011.

al. (1997); Marriott and Faber-Langendoen (2000); Walford et

38. Cooper and Andrus 1994.

al. (2001); and reports accessible at the Wyoming Natural Diver-

39. Mitsch and Gosselink 2007; Copeland 2010; Keddy 2010.

sity Database website (http://www.uwyo.edu/wyndd/reports

40. The American Society of Wetland Managers estimates a loss of

-and-publications/), authored by W. Fertig, B. Heidel, M.

38 percent of Wyomings original 2 million acres of wetlands.

Jankovsky-Jones, G. Jones, and others.

Gage and Cooper (2012) state that 3540 percent of the wet-

19. Salt flats are a common feature in desert shrublands; see chapter
8.

lands present in Wyoming, Nebraska, and South Dakota at the


end of the eighteenth century have been lost, but they give no

20. The salt deposits are commonly referred to as alkali and the

estimates for each state. Both estimates include nonriparian

salt flats as alkali flats. However, alkalinity is related to acid-

and riparian wetlands. Dahl (1990) estimates 38 percent wet-

ity, not to the concentration of salts (carbonates, bicarbonates,

land loss for Wyoming between 1780 and the mid-1980s.

sulfates, and chlorides of calcium, magnesium, potassium, and

41. Youngblood et al. 1985; Hansen 1995; Girard et al. 1997.

sodium); Hart and Lovvorn 2000; Gage and Cooper 2012.

42. Gage and Cooper 2012.

21. Holpp 1977; Wollheim and Lovvorn 1995; Brough 1996; Bartz
1997; Bowman 1997; Sanderson et al. 2008; reports available
from the Wyoming Natural Diversity Database website (http://
www.uwyo.edu/wyndd/reports-and-publications/).
22. Keddy 2010.
23.
See reports available at the Wyoming Natural Diversity
Database website (http://www.uwyo.edu/wyndd/reports-and
-publications/).

43. Zedler and Kercher 2004; Zedler 2011.


44. Garrison creeping foxtail has similar effects, though it is introduced; see chapter 17.
45. Kulmatiski et al. 2010; Gage and Cooper 2012; Uddin et al.
2012.
46. Zedler and Kercher 2004; Jarchow and Cook 2009.
47. The Flora of North America (volume 22, 2000) indicates that
narrowleaf cattail has expanded its range and has become more

Notes to Pages 8094


abundant, especially in roadside ditches and other highly dis-

9. Nematodes are a highly diverse group of microscopic round-

turbed habitats. The hybrid (Typha glauca) tends to occur in

worms. Some grassland species feed on plant roots; others feed

such places and is more invasive than narrowleaf cattail.

on fungi, bacteria, or protozoans. Some fungi have evolved

48. Poiani and Johnson 1991, 1993; Johnson et al. 2010; Shinker et
al. 2010.
49. Dise 2009; the dynamics of peatlands are complicated, and the
consequences of climate change are difficult to predict.

adaptations for trapping and digesting nematodes. Stanton


1988; Moore et al. 2008.
10. Weaver 1958,1968; Risser 1985.
11. Nippert and Knapp 2007.

50. Shinker et al. 2010.

12. Singh et al. 1983b; Lauenroth et al. 2008.

51. Copeland et al. 2010.

13. Ode and Tieszen 1980; Risser 1985.

52. Copeland et al. 2010; Sueltenfuss et al. 2013.

14. Boutton et al. (1980) study an elevational gradient from the top

53. Peck and Lovvorn 2001; data were obtained during a relatively
wet year.

of the Laramie Mountains to near Cheyenne.


15. Peden 1976; Vavra et al. 1977; Schwartz and Ellis 1981.

54. Copeland et al. (2010) review threats to the integrity of wetlands

16. Coupland and Van Dyne 1979; the weight of all microbial

in Wyoming. For a summary of federal regulations pertaining to

organisms in the soil is about half that of aboveground plant

wetland protection, see the report by Claudia Copeland (2010).

biomass in a Canadian mixed-grass prairie (Paul et al. 1979).

Also, see the summary prepared by the Colorado Natural Heritage

Stanton (1988) provides data on the relative abundances of

Program at http://www.cnhp.colostate.edu/cwic/regulations.asp.

different kinds of microorganisms. Moore et al. (2008) list

55. National Academy of Sciences 2001; Mitsch and Gosselink 2007

the following functional groups in the belowground food

(chapter 12); Sueltenfuss et al. 2013.

web: predatory mites, nematophagous mites, predatory nematodes, fungivorous nematodes, bacteriophagous nematodes,

Chapter 6: Grasslands

collem
bolas, mycophagous prostigmata, amoebae, flagellates,
and phytophagous nematodes.

1. In Wyoming, buffalo grass is found most commonly in the

17. Sims and Coupland 1979; Singh et al. 1983b; Lauenroth et al.

shortgrass prairie south and east of Cheyenne and sporadically

2008; aboveground plant growth is between 100 and 300 g/m 2/

into Montana. For more information on grassland distribution,

year, but belowground estimates are problematic because of the

see Singh et al. (1983a,b). Lauenroth and Burke (2008) review


the ecology of shortgrass prairie and classify most Wyoming
grasslands as northern mixed prairie.
2. There are occasional small patches of tallgrass prairie on sandy
soils and in some riparian zones on the eastern plains; common plants include big bluestem, Indiangrass, prairie dropseed,

difficulties of measuring root growth.


18. Sala et al. 1988.
19. Sala et al. 1988; Oesterheld et al. 1999; Lauenroth et al. 2008.
20. Coughenour 1991a,b.
21. Knapp and Seastedt 1986; Hunt et al. 1987; Lauenroth and
Burke 2008.

sideoats grama, and switchgrass. Buffalo grass reproduces veg-

22. Milchunas et al. 1988, 2008; Stohlgren et al. 1999.

etatively by stolons, which is unusual in the grass family.

23. Ingham et al. 1985; Knapp and Seastedt 1986; Elliott and Cole-

3. Forbs are herbaceous plants other than grasses and sedges

man 1988; Stanton 1988; Holland and Detling 1990; Burke et al.

(graminoids). Shrubs produce woody stems above the soil sur-

1997, 2008; Frank and Evans 1997; Oesterheld et al. 1999; Frank

face. Very small shrubs, such as fringed sagewort, are sometimes

et al. 2000, 2011.

referred to as half-shrubs. Forbs have more species in grasslands

24. Johnson and Rumbaugh 1981.

than any other growth form.

25. West 1990.

4. Epstein et al. 2002; Morgan et al. 2007, 2011.

26. Woodmansee et al. 1978.

5.
Hudson 1986; see also http://news.sciencemag.org/plants

27. Lauenroth et al. 1978; Lauenroth and Dodd 1978, 1979; Risser

-animals/2013/03/fairy-circle-mystery-solved.
6. Lauenroth et al. (2008) review above- and belowground biomass and productivity in grasslands.
7. Bud dormancy can be broken by changes in the production of

and Parton 1982; Wight and Godfrey 1985; Lauenroth and


Burke 2008; Dijkstra et al. 2010; Carrillo et al. 2012.
28. Houston et al. 1973; Houston and Hyder 1975; Rauzi and Fairbourn 1983; Senft et al. 1985.

growth-stimulating hormones, which often occur when leaves

29. Clark 1977 (p. 1322); Clark et al. 1980; Polley and Detling 1988.

and stems are eaten or are damaged in some other way. The

30. Burke et al. 1987.

buds remain dormant as long as the leaves and stems are intact,

31. Succession following the abandonment of agricultural land is

a process known as apical dominance. Briske (1991) reviews grass

described by Lang (1973) and Reichardt (1982); Peters et al.

morphology and physiology as it pertains to herbivory.

(2008) provide an overview of disturbances in grasslands,

8. Richards and Caldwell (1985) found that most of the energy

including the effects of fecal deposits from large mammals

required for regrowth in two species of wheatgrass was derived

and mortality caused by white grubs (the larvae of June

from current photosynthate by remaining leaves rather than

beetles).

from soluble carbohydrates stored in roots. Lauenroth et al.

32. Painter and Belsky 1993; Lauenroth et al. 1994; Oesterheld et

(2008) report that belowground plant growth is about three

al. 1999; Stohlgren et al. 1999; Lauenroth and Burke 2008; Mil

times more rapid than aboveground in shortgrass prairie.

chunas et al. 2008.

331

332

Notes to Pages 95102


33. Clipping plants uniformly may not give the same results as

57. Wright and Bailey 1980 (for the Great Plains); Bragg 1982 (for

actual grazingwhich typically leaves some part of the plant

central Nebraska); Wendtland and Dodd 1992 (for western

uneaten or severed at different heights. The novel idea that

Nebraska).

plant growth-promoting substances might occur in the saliva

58. Knapp and Seastedt 1986.

of mammalian herbivores has not been supported by most

59. Redman (1978) found more water stress and lower plant growth

research (Detling and Dyer 1981; Detling et al. 1981), although

a year after burning mixed-grass prairie in Saskatchewan, and

sheep saliva may enhance short-term growth under certain

Engle and Bultsma (1984) suggest that increases in plant growth

conditions (Gullap et al. 2011).

should not be expected following burning during drought years

34. Owen and Wiegert 1976, 1981, 1982; Briske 1991.

in western South Dakota. See also Scheintaub et al. 2009.

35. Painter and Detling 1981; Detling and Painter 1983.

60. Fuhlendorf et al. 2012; McGranahan et al. 2012.

36. McNaughton 1985 (p. 286).

61. Branson and Miller 1981; Branson 1985.

37. Heavy grazing by bison has been discussed by Larson (1940),

62. The effects of drought can vary; several investigators reported

Roe (1951), Lauenroth et al. (1994), and Hart (2001). McGinnies

that blue grama and buffalo grass increase in abundance with

et al. (1991) report on the resiliency to drought of grasslands in

extended drought (Albertson and Weaver 1944; Newbauer et al.

eastern Colorado; Laycock (1991) and Cid et al. (1991) review

1980), even to the point of creating a shortgrass prairie. Others

the conditions that favor recovery.

have reported a decline of the same species during similar con-

38. Davis 1959 (p. 4); Sturgis 1884 (p. 6).


39. Quoted in Branson 1985 (p. 4); April 20, 1805, in the Meriwether Lewis journal, published in 1814.
40. Houston 1963; Lauenroth et al. 2009.
41. Van Vuren (1987) and Urness (1989) did not agree with Mack

ditions (in southeast Montana: Ellison and Woolfolk 1937; Reed


and Peterson 1961).
63. Reed and Peterson 1961; Branson and Miller 1981.
64. Turner and Costello 1942; Houston 1963.
65. Weaver 1943, 1954; Newbauer et al. 1980.

and Thompsons (1982) explanation, suggesting instead that

66. Reed and Peterson 1961; Rodel 1977; Watts et al. 1982; the

bison numbers were sometimes low in the Great Basin because

distribution and characteristics of different species of grass

of intensive hunting by Indians. Early journals suggest large

hoppers in Wyoming are summarized by Lockwood et al.

numbers of bison in western Wyoming (L. Whittlesley, pers.

(1993) and Crist (2008). There are approximately 100 species of

comm.; Hart 2001).


42. K. Johnson (1987) includes historical photos; the relative effect
of climate and grazing is discussed by Milchunas et al. (1989).

grasshoppers in the state.


67. Watts et al. 1982 (p. 288).
68. Ellison 1960; Holmes et al. 1979; Quinn and Walgenbach 1990.

43. Weaver and Clements 1938 (p. 470).

69. Rodell 1977; Hewitt and Onsager 1983.

44. Ellison 1960 (p. 52).

70. Allred 1941 (p. 387).

45. Rauzi and Smith (1973) and Gifford and Hawkins (1978) review

71. The quote is from Pfadt and Hardy (1987, p. 183); Crist (2008)

the effects of livestock grazing on water infiltration.


46. Coughenour 1991a,b.
47. Briske et al. 2005, 2006, 2011.
48. McNaughton 1976; several studies have found that the growth
forms of western wheatgrass on prairie dog towns were shorter
and more prostrate than those away from the towns, suggest-

states that the effects of livestock grazing on grasshopper abundance are still uncertain, and the number of insects may be
triggered more by weather and plant growth.
72. Lockwood et al. 2000, 2002; Zimmerman et al. 2004.
73. Dorn 1986; Lockwood 2004; locusts are a form of grasshopper;
Mormon crickets are katydids, not crickets.

ing that some varieties of a single species tolerate grazing bet-

74. Foster and Stubbendieck 1980; Laycock (1958) studies the re

ter than do others; see Detling and Painter 1983; Detling et al.

vegetation of pocket gopher mounds; Peters et al. (2008) and

1986; Jaramillo and Detling 1988; Whicker and Detling 1998.

Stapp et al. (2008) review the effects of small mammals and

49. Coppock and Detling 1986; Krueger 1986; Day and Detling
1990a,b; Fahnestock and Detling 2002; Johnson-Nistler et al.
2004.

insects.
75. Archer et al. 1987; Whicker and Detling 1998; Sharps and Uresk
1990. Peters et al. (2008) report that prairie dog burrows may

50. Mielke (1977) suggests a mutualistic relationship between bison

be 39 feet deep, with 50300 mounds or more per colony.

and pocket gophers, with the bison improving conditions for

Towns can be tens to a thousand acres or more in size. Stapp

the forbs preferred by the gophers. The pocket gophers, he sug-

et al. (2008) and Davidson et al. (2012) provide an overview of

gests, improve soil fertility by their burrowing and feeding,


thereby increasing forage production for the bison.

burrowing mammals in grasslands.


76. Densmore (2011) reviews the history of black-footed ferrets,

51. Kamm et al. 1978; Ingham and Detling 1984.

which were declared extinct in 1979. Two years later, a colony

52. Ingham et al. 1985.

of about 130 animals was found near Meeteetse in the Bighorn

53. Oesterheld et al. 1999.

Basin. That number was soon reduced to 18 by canine distemper

54. Higgins 1984.

and sylvatic plague, for which the animals are now vaccinated

55. Woody plants on the fringe of burning grasslands are some-

when possible. After capture and a period of breeding in captiv-

times ignited and could sustain a fire during a thunderstorm.


56. Moore 1972.

ity, ferrets have been released where large prairie dog colonies
can be found (with about 1,000 ferrets now living in the wild

Notes to Pages 103116


and 300 still in captivity). One of the first reintroduction sites

95. Morgan et al. 2011.

was the Shirley Basin north of Medicine Bow, where a breed-

96. Morgan et al. 2011 (p. 202).

ing population of several hundred ferrets has been established.

97. Dijkstra et al 2010; Carrillo et al. 2012; Nie et al. 2012.

Each ferret eats about 100 prairie dogs a year. Most present-day

98. Epstein et al. 2002; Morgan et al. 2007.

prairie dog towns are too small to sustain the ferret (Davidson
et al. 2012).
77. Freilich et al. 2003; Miller et al. 2007; Davidson et al. 2012.
78. Stapp et al. (2008) suggest that cultivation has more adverse
effects than shooting or poisoning and that conservation efforts

Chapter 7: Sagebrush
1. Driese et al. 1997.
2. Except in the southeast corner, the grasslands of Wyoming have

could be more fruitful by protecting land where the prairie dogs

been classified as northern mixed-grass prairie; the term steppe

are likely to establish new colonies, such as abandoned prairie

implies large areas of grassland or shrubland, often extending

dog towns. It is important to protect large colonies, as they are


a source of dispersing animals for other towns, and also because
larger towns provide more benefits to the associated species.
79. Pauli et al. 2006.
80. Scott 1951; Kirkham and Fisser 1972; Sneva 1979; Rogers 1987;
Peters et al. 2008.

to the horizon.
3. Houston 1961; West 1983, 1988; Miller et al. 1993; Driese et al.
1997; Bates et al. 2006; Poore et al. 2009; Schlaepfer et al. 2011.
4. Beetle and Johnson 1982; Barker and McKell 1986; Kolb and
Sperry 1999a,b.
5. Tall plants of basin big sagebrush are found in the dunes north-

81. Allen and Knight 1984; Evans et al. 2001; Ehrenfeld 2003;

east of Rock Springs; in the Ferris Dunes north of Rawlins; and

Strayer et al. 2006; Bradley and Marvin 2011. The adaptations

at the base of impermeable granitic domes, where water runs off

of invasive plants can include the production of large volumes

into the adjacent soil.

of seed that germinate earlier in the spring than the seeds of

6. Winward 1970; Shumar and Anderson 1986.

native species, and chemical defense compounds that local her-

7. Sand sage (Artemisia filifolia) and threetip sagewort (A. tripartita)

bivores and parasites cannot tolerate (Pfister et al. 2002).

also are capable of sprouting.

82. Bradley and Marvin 2011; in addition to cheatgrass, drought-

8. Welsh 1957; Robertson et al. 1966; Smith 1966; Myers 1969;

tolerant invasives that could become more common in Wyo-

Zamora and Tueller 1973; Sabinske and Knight 1978; Baker and

ming grasslands include camelthorn, medusahead wildrye,

Kennedy 1985. West of Laramie, black sagebrush is found on

rush skeletonweed, and ventenata grass.

soils with a hardpan at 812 inches, sometimes intermingled

83. Blossey and Notzold (1995) propose the Evolution of Increased


Competitive Ability hypothesis.
84. Strayer et al. 2006.
85. In Nevada: Leger 2008; Goergen et al. 2011; Rowe and Leger
2011; in Wyoming: Mealor and Hild 2007.
86. Blumenthal et al. 2003, 2010; Perry et al. 2010.
87. Blumenthal et al. (2008) added snow water to one plot of

with patches of grassland on more shallow soils.


9. Burke et al. 1989.
10. Lundberg 1977; a desert pavement develops when strong winds
remove fine soil particles from the surface, leaving a layer of
stones too heavy for the wind to move.
11. West and Klemmedson 1978; Allen and MacMahon 1985; Longland and Bateman 2002.

mixed-grass prairie and observed an increase in the abundance

12. West and Young 2000.

of invasive plants.

13. Richards and Caldwell 1985; Caldwell et al. 1998; Ryel et al.

88. Miller et al. 2007; see note 76.

2002, 2004; Leffler et al. 2004, 2005; Meinzer et al. 2004;

89. Wyeth 1899; Carter 1857, as reported in Carter 1939.

Scott et al. 2008; Aanderud and Richards 2009; Bleby et al.

90. Roach et al. (2001) discuss the genetic implications of small,

2010; Prieto et al 2011; see also http://en.wikipedia.org/wiki/

isolated prairie dog colonies, finding that in some areas disper-

Hydraulic_redistribution.

sal from one colony to another is good if there is a network

14. Naithani et al. 2012.

of swales and drainages along which dispersing animals can

15. Richards and Donovan 2005; Naithani et al. 2012.

travel; further isolation could lead to inbreeding and a reduc-

16. DePuit and Caldwell 1973.

tion in genetic diversity.

17. Beetle and Johnson 1982; Miller and Schultz 1987.

91. Grassland National Park in Saskatchewan was established in

18. Welch et al. 1991. Water extracts and volatile vapors from big

1981; in 2009 it was designated as the worlds largest Dark Sky

sagebrush leaves are thought to affect plant growth (Hoff-

Preserve.

man and Hazlett 1977; Weaver and Klarich 1977; Ferreira and

92. Freilich et al. (2003) describe how livestock production has

Janick 2004), but until field experiments are conducted, it

changed grassland ecosystems and some of the steps that might

is prudent to assume that the primary adaptation of volatile

be taken to restore some lost features.

oils is to reduce herbivory rather than retard the growth rates

93. Wiens and McIntyre (2008) note that populations of grassland

of neighboring plants. Volatile sagebrush compounds may

birds are declining more than any other group of North Ameri-

reduce the rate of nitrification (West 1991); some evidence

can birds.

suggests that a browsed shrub emits volatile cues that reduce

94. Burke et al. (1995) estimate 50 years for recovery of soil organic
matter.

herbivory (insect or ungulate) on adjacent plants (Karban et al.


2011).

333

334

Notes to Pages 116125


19. Some herbivores have evolved behavioral and physiological

46. Lommasson 1948; Frischknecht and Plummer 1955; Mueggler

adaptations for minimizing the adverse effects of terpenes and

1956; Robertson et al. 1966; Johnson 1969.

other plant toxins (Launchbaugh et al. 2001); see also Pfister et

47. Lesica et al. 2007; Ewers and Pendall 2008.

al. 2002.

48. Cooper 1953; Anderson and Holte 1981; Uhlich 1982; Anderson

20. Passey and Hugie 1963; Ferguson 1964; Biondi et al. 2007.
21. Robertson et al. 1966; Choudhuri 1968; McDonough and

and Inouye 2001; Ewers and Pendall 2008.


49. Lang 1973; in the Great Basin, Morris et al. (2011) found that

Harniss 1974; Daubenmire 1975; West et al. 1979; Cawker

the effects of cultivation persisted for more than a century.

1980; Romo 1984; Eckert et al. 1986; Meyer and Monsen 1991;

50. Allen et al. 1987; Stahl et al. 1998; Frost et al. 2001; Schuman et

Schuman et al. 2005.

al. 2005; Hild et al. 2006; Boyd and Svejcar 2011.

22. Choudhuri 1968.

51. Ellison and Woolfolk 1937.

23. Lommasson 1948; Robertson et al. 1966; Schuman et al. 2005.

52. Nelson and Sturges 1986; Allen et al. 1987.

Late spring frosts and cold winter temperatures can be fatal to

53. Hanson et al. 1982; Nelson and Tiernan 1983.

seedlings; the development of mycorrhizae on sagebrush roots

54. West 1988; Wallace and Nelson 1990.

facilitates establishment (Allen et al. 1987; Stahl et al. 1998;

55. Neilson et al. 2005; Poore et al. 2009; Finch 2012.

Schuman et al. 2005).

56. Allred 1941 (p. 387).

24. Daubenmire 1975; West et al. 1979; Cawker 1980.

57. Cowan 1929; Allred 1941; Furniss and Krebill 1972; Parmenter

25. Schuman et al. 2005; Hild et al. 2006.

et al. 1987; Miller et al. 1993. Sagebrush is commonly infested

26. Sturges 1975.

with gall-forming flies, but apparently with little detrimental

27. Harniss and Murray 1973; West 1983; Singh et al. 1983b.

effect; the effects of voles are described by Frischknecht and

28. Robertson 1947; Sonder 1959.

Baker (1972).

29. Inouye 2006.

58. Zimmerman et al. 2004; see also chapter 6.

30. Sheets 1958; Tabler 1968; Sturges 1986.

59. Homsher 1960 (p. 192).

31. Grayson 2006; Mack and Thompson 1982; Van Vuren 1987;

60. Miller et al. 1993; Sturges 1993; Mensing et al. 2006; Morris et

Urness 1989; Miller et al. 1993; bison abundance in the inter-

al. 2010; Nelson and Pierce 2010.

mountain basins may have varied from place to place, depend-

61. Baker 2006, 2011.

ing on the amount of grassland available. Large herds of bison

62. Wright et al. 1979.

were noted frequently by explorers traveling through western

63. Young and Evans 1978; Wright et al. 1979; Rennick 1981.

Wyoming in the early 1800s (L. Whittlesley, pers. comm.); see

64. Ewers and Pendall 2008; Cleary et al. 2010.

the journals of James Clyman, Warren Ferris, Zenas Leonard,

65. Cheatgrass, common dandelion, and yellow salsify; Wright et

W. F. Raynolds, Osborne Russell, Robert Stuart, and Nathaniel


Wyeth, summarized in Dorn (1986).
32. Mehringer and Wigand (1990) found that, over the past 5,500

al. 1979.
66. Mihlbachler 1986.
67. Smith et al. 1985; Baker 2006.

years, the abundance of charcoal has been inversely related

68. Chambers et al. 2007; Mack 2011a,b; Rau et al. 2011.

to the abundance of sagebrush pollen and directly related to

69. Bradley 2010; Bradley and Marvin 2011.

the abundance of grass pollen in the Great Basin of Utah. This

70. Tyrer et al. 2007; rush skeletonweed is now found in Idaho.

observation suggests that grasslands burned more often than

71. Young and Allen 1997; West and Young 2000.

sagebrush steppe.

72. Seefeldt et al. 2007; Davies et al. 2009b; Shinneman and Baker

33. Baker 2011.


34. Pearson 1965a.
35. Joyce 1981.
36. Mummey et al. 2002; Bechtold and Inouye 2007; Davies et al.
2007a,b, 2009a,b; Kwon et al. 2008; Wick et al. 2009.
37. Huxman et al. 2004; Prater et al. 2006; Kwon et al. 2008; Cleary
et al. 2010.
38. Havstad et al. 2007; Cleary et al. 2010.

2009; Prevey et al. 2010a,b.


73. The soil may have sufficient cheatgrass seedthe soils seed
bankto produce more than 2,000 seedlings per square yard
(Young and Evans 1975; Hassan and West 1986).
74. Young et al. 1976; Whisenant 1990; forage availability from one
year to the next can be highly variable where annual plants
dominate the vegetation, which creates problems for livestock
and wildlife management.

39. Murray 1975; Carpenter and West 1987; for nitrogen limita-

75. Gill and Burke 1999; Norton et al. 2004; Prater et al. 2006; John-

tions, see Charley (1977) and Doescher et al. (1990); West and

son et al. 2011; Mack 2011a,b; Rau et al. 2011; see Ostoja and

Skujins (1977) observed nitrogen fixation in microphytic crusts.

Schupp (2009) for data on small mammals, which are less com-

40. Matson et al. 1991.


41. Branson et al. 1976; Parmenter et al. 1987.

mon in cheatgrass than in adjacent big sagebrush.


76. Shinneman and Baker 2009; Prevey et al. 2010a,b; Eiseroad and

42. Charley and West 1975; Charley 1977.

Rudd 2011; Owen et al. 2011; Meyer et al. (2007) describe a fun-

43. Burke et al. 1989.

gal pathogen that may have possibilities for biocontrol; see also

44. Friese and Allen 1993; Soul and Knapp 1996; Eldridge 2004,
2009.
45. McCulley et al. 2004.

chapter 6.
77. Havstad et al. 2007.
78. Vale 1975; Dorn 1986; Johnson 1987; Miller et al. 1993.

Notes to Pages 125135


79. Fremont 1845 (p. 49).
80. Preuss 1958 (p. 23); Platt and Slater 1852 (p. 12).

3. Smith and Nowak 1990; West (1983) reports that saltbushgreasewood desert shrublands have a mean net aboveground

81. Bradley (2009) identified the Powder River Basin as an area

primary productivity of 20100 g/m2/year and big sagebrush

where big sagebrush ecosystems are at relatively high risk due

steppe a mean of 80250 g/m2/year. Carbon fixation by biotic

to climate change and other factors.


82. Reed and Peterson 1961.

soil crusts can be considerable when the soil surface is wet.


4. The most common salts in Wyoming desert shrublands and

83. Great year-to-year variation in shadscale-dominated desert

playas are magnesium and sodium sulfates, sodium chloride,

shrubland has been documented in the Great Basin through

and bicarbonates of calcium and magnesium. Less common

photographs taken of the same area annually by Sharp et al.

are calcium and magnesium sulfite salts and calcium sulfate

(1990).

di
hydrate (gypsum). All soils with high salt concentrations

84. Kleiner and Harper 1977; West and Young 2000; Belnap and

can be referred to as saline, with the term alkaline used for soils

Lange 2001; Muscha and Hild (2006) study soil crusts in Wyo-

where the hydroxides cause a pH of 8.4 or above. The term

ming; Shinneman and Baker (2009) found more rapid cheat-

sodic is used for soils where 15 percent or more of the exchange

grass invasion on sites with low cover of soil crusts.

capacity is occupied by sodium. Sodic alkaline soils commonly

85. Mack and Thompson 1982.

have high pH (exceeding 9 in some cases) and low rates of water

86. Van Vuren (1987), Urness (1989), and Byers and Smith (2007)

infiltration and air permeability. The soils of greasewood tend

emphasize the importance of intensive hunting by Indians; see

to be alkaline with a pH of 8 or higher and a conductivity of

also Grayson 2006.


87. About James Clyman: Wyeth 1899; Lee (1916, p. 134); Camp

410 dS/m or greater.


5. The adaptations of halophytes and other desert plants are

1960 (p. 28). About William P. Hunt: Morris 2013 (p. 123);

described by Jefferies (1981), Smith and Nowak (1990), Gure-

Townsend 1839; see also Dorn 1986.

vitch et al. (2002), and Sperry and Hacke (2002).

88. Donahue 2010.

6. Goodin and Mozafar 1972; Detling and Klikoff 1973.

89. Manier and Hobbs 2007.

7. White 1976.

90. Beever 2003; Gardner 2008; De Steigure 2011; Philipps 2012.

8. West 1983.

91. See http://www.blm.gov/wy/st/en/programs/Wild_Horses.html.

9. Nichols 1964; Sharma and Tongway 1973; Romo 1984; Romo

92. See http://gf.state.wy.us/wildlife/nongame/LIP/Sagebrush/index


.asp; see also Welch 2005; Holmes 2007.
93. Baker 2006, 2011 (on fire); Neilson et al. 2005; Madson 2006a,b;

and Eddleman 1985; Ladenburger et al. 2006.


10. Khan et al. (2002) found that greasewood seed germination
declines with increasing salinity.

Meinke et al. 2009; Davies et al. 2011 (on energy development);

11. Hamner 1964; Nichols 1964.

Poore et al. 2009 (on climate change); see chapter 18 for discus-

12. Nichols 1964; Brown 1971.

sion of habitat fragmentation.

13. Flowers and Evans 1966; Steger 1970; Detling and Klikoff 1973;

94. Crawford et al. 2004; Gilbert and Chalfoun 2011; see also http://
sagegrouseinitiative.com/.
95. The sharptailed and dusky grouse also use sagebrush in some
places.
96. Fedy et al. 2011.

Skougard and Brotherson 1979.


14. McGinnies et al. 1976; Bowman et al. 1985; Blaisdell and Holm
gren 1984; Tiku (1976a,b) and Cluff et al. (1983) found that saltgrass tolerates soils with a conductivity ranging from 5 to 25
dS/m and a pH of 710.

97. Blickley et al. 2012.

15. Branson et al. 1976; Baker and Kennedy 1985.

98. Dinkins et al. 2012.

16. Gates et al. 1956; Hamner 1964; big sagebrush has been reported

99.Baker 2006, 2011; Beck et al. 2009; Coates and Delehanty 2010;
Doherty et al. 2011a,b; Hess and Beck 2012a,b.
100.Holloran et al. 2005; Beck et al. 2012; Hess and Beck 2012b;
Kirol et al. 2012.
101.Naugle et al. 2004, 2005; Walker et al. 2007; Knick and Con-

on soils with a pH of 8.4 and a relatively low conductivity of


3.4dS/m.
17. Blaisdell and Holmgren 1984.
18. Beath et al. 1941; Kolm 1975. There are two groups of plants
that accumulate selenium. One group, the selenium indica-

nelly 2011; Dinkens et al. 2012.

tors, requires selenium for growth and includes two-grooved

102.Neilson et al. 2005; Poore et al. 2009.

milkvetch, tineleaved milkvetch, woody aster, and princes

103.Madson 2006a; Doherty et al. 2011a,b; Bush et al. 2011; Dzialak et al. 2011; Knick et al. 2011; Reese and Connelly 2011.

plume. In contrast, the other comprises facultative accumulators, such as Gardners saltbush, gumweed, locoweed,
mentzelia, snakeweed, spiny hopsage, winterfat, and certain

Chapter 8: Desert Shrublands and Playas

grasses. Plants in the latter group do not require selenium


but often develop high concentrations of the element. Some

1. West and Young 2000.

plants do not accumulate selenium even though they are fre-

2. Drenovsky and Richards 2006; Drenovsky et al. 2010; some

quently found on seleniferous soils. Both selenium indicators

shrubs are clearly deciduous, such as shadscale and greasewood;

and facultative accumulators can concentrate large amounts

others are evergreen or facultatively deciduous, such as fourwing

of the element in leaves and stems without adverse effects on

saltbush, Gardners saltbush, spiny hopsage, and winterfat.

the plant itself. For example, two-grooved milkvetch is known

335

336

Notes to Pages 135148


to accumulate 4,000 ppm or more from soils containing only
1.1 ppm selenium. Grasses and cereal grains with only 1030

Desert shrubs probably would have increased in density during


this time, reducing the growth of the preferred forage plants.

ppm can be lethal or crippling to cattle and horses. Both the

37. Chambers and Norton 1993; Young 1994.

indicators and the accumulators increase surface-soil selenium

38. West and Young 2000; Grayson (2006) concluded that there

concentrations as their shoots decompose in the fall, thereby

would have been large grazers in the late Pleistocene (as discussed

possibly improving conditions for their own growth.

in chapter 2), but for tens of thousands of years, the plants of

19. Harris 1992.


20. Hamner 1964; Hodgkinson 1987.

most intermountain basins probably underwent relatively light


grazing by large mammals, compared to the western Great Plains.

21. Russey 1967; Gibbens 1972; Blaisdell and Holmgren 1984;

39. Muscha and Hild (2006) studied soil crusts in Wyoming;

several studies have found that saltbush shrubland exists on

Shinneman and Baker (2009) report rapid cheatgrass invasion

various soil textures, including sandy loam, sandy clay loam,

where the cover of soil crusts is relatively low.

loam, and clay loam. The soils tend to have low infiltration rates

40. Meyer et al. 2001.

compared to where big sagebrush and blue grama are common,

41. Scott et al. 2010.

probably because of their relatively high clay content or the dis-

42. Shinneman and Baker 2009.

persing effect of sodium on surface-soil particles. A soil pH of 8

43. Rickard 1964.

or above is common. The soils are not always saline, sometimes


having a conductivity of 1.5 dS/m or less. Soil salinity is often
higher under shrubs than between them.

Chapter 9: Sand Dunes, Badlands, Mud Springs, and Mima Mounds

22. Stevens and Monsen 1988.

1. Ahlbrandt 1973; Mayer 2003.

23. Branson et al. 1976.

2. Gaylord 1982.

24. Harr and Price 1972.

3. Prairie sandreed and blowout grass are restricted to dunes in the

25. Rickard 1967.


26. Meyer et al. 2001.
27. West 1991; James et al. (2005) suggest that phosphorus and
magnesium also can be limiting for salt desert shrub growth.
28. West 1990, 1991; Miller et al. 1993; West and Young 2000; Belnap and Lange 2001; nitrogen fixation by soil crusts may be bal-

eastern half of Wyoming, perhaps because of their intolerance


of the drier dunes to the west.
4. Barnes and Harrison 1982.
5. Chadwick and Dalke 1965.
6. Shumway 2000.
7. Wolfe (1973) studied the effects of fire on stabilized dunes;

anced by nitrogen losses caused by denitrification when the soil

chapter 2 of this book reviews the effects of drought.

is wet and anaerobic; for a discussion of biological soil crusts,

8. Weaver and Albertson 1956; Bleed and Flowerday 1990.

see www.soilcrust.org.

9. Pearson 1965b.

29. Currie and Goodwin 1966; Plummer et al. 1968.

10. High infiltration and storage in the Nebraska Sand Hills provide

30. Sharpe and Van Horne 1998.

much of the groundwater contained in the Ogallala aquifer, on

31. Wight and Nichols 1966; Kirkham and Fisser 1972.

which many Great Plains municipalities and farmers depend

32. Allred 1941; Blaisdell and Holmgren 1984.

(Zwingle 1993); see Ashworth (2006) for a modern synthesis of

33. Sharp et al. 1990.

the geologic development and status of the Ogallala aquifer.

34. Wallace and Nelson 1990.

11. Bleed and Flowerday 1990.

35. Succession in desert shrublands subjected to livestock grazing

12. Koske and Polson 1984; Brown 1997.

has been studied in several locations. Predictably, the most pal-

13. Barnes and Harrison 1982; Potvin and Harrison 1984.

atable species tend to decrease in abundance, for example, four-

14. Nitrogen fixation also occurs in the rhizosheaths of some

wing saltbush, spiny hopsage, winterfat, and some grasses. Less

grasses, including Indian ricegrass, prairie sandreed, and

palatable species increase, such as rabbitbrush, shadscale, and

needle-and-thread grass (Bergmann et al. 2009).

weedy species (including Russian thistle). The season of graz-

15. Martin 1992; Clarke and Rendell 2003; Forman and Pierson

ing can be important in determining rangeland response, with

2003; Mayer and Mahan 2004; Kelly et al. 2008; Hanson et al.

early-winter sheep grazing generally having less impact on desirable species than late-winter or early-spring grazing. Ranchers

2009, 2010; Schmeisser et al. 2009; Halfen et al. 2010.


16. Yizhaq et al. 2009.

commonly put sheep on the desert in winter and then transport

17. Sahara mustard is a problem in the dunes of California.

them to mountain rangelands in the summer. With protec-

18. Brown 1971; the Powder River Breaks are found on weakly

tion from livestock grazing, palatable species (perennial grasses

consolidated shales interspersed with lignite seams in the Fort

and forbs) increase, and the abundance of introduced annuals

Union Formation, from which most Wyoming coal is mined.

may decrease or remain about the same. Rangeland recovery in

Vegetation banding caused by abrupt changes in bedrock is

exclosures is variable, which has led to the hypothesis that some

found in various places in Wyoming, such as near Lander and

ecosystems are permanently changed after a disturbance of sufficient intensity and duration (the state transition hypothesis).
36. West and Young (2000) present evidence that the grazing
capacity of desert shrublands had been exceeded by 1900.

Kemmerer.
19. Similar mounds are found in northern Albany County, north of
Laramie (L. Munn, pers. comm.).
20. Both quotes are from Hayden (1879, p. 130).

Notes to Pages 149173


21. Cox 1984, 2012; Washburn 1988.

22. Betancourt et al. 1991; pinyon pine populations also occur in

22. McGinnies 1960; Hansen 1962.

Idaho (west of Kemmerer) and in northern Colorado (Owl Can-

23. L. Munn, pers. comm.

yon, north of Fort Collins, east of Livermore).

24. Mears 1981, 1987.


25. Spackman and Munn 1984.
26. McFaul 1979.

23. Progulske 1974; K. L. Johnson 1987; Veblen and Lorenz 1991;


see note 27 in chapter 11 for the subtle distinction between fire
exclusion and fire suppression.

27. Reider et al. 1996, 1999.

24. Orr 1975; Shepperd and Battaglia 2002.

28. Arkley and Brown 1954 (p. 198).

25. Tolstead 1947; Potter and Green 1964; Weaver 1965; Brown 1971;

29. Arkley and Brown 1954; Cox 1984 (p. 45) Cox and Allen 1987;
Cox et al. 1987; Washburn 1988; Cox and Hunt 1990; Gabet et
al. 2014.
30. Cox 1990.
31. Burnham and Johnson 2012. The quotation is from Berg 1990
(p. 284).

Hoffman and Alexander 1987; Hansen and Hoffman 1988.


26. Welsh 1957; possibly, ponderosa pine can survive in western
Wyoming, but it has not yet migrated to the suitable habitats
there (see chapter 2).
27. States 1968.
28. Ponderosa pine is found with Rocky Mountain juniper on outcrops of shale that are rich in bentonite, such as near Upton

Chapter 10: Escarpments and the Foothill Transition


1. Kornfeld and Osborn (2003) discuss the ecological and sociological effects of patchy landscapes.

and Moorcroft; elsewhere it is found on outcrops of sandstone


(for example, southwest of Casper and north of Lusk) or granite (for example, west of Cheyenne and Wheatland).
29. Lepper 1974.

2. Hanna 1934; Myers 1969; see chapter 17.

30. Despain 1973, 1990.

3. States 1968; Wells 1970.

31. Gruell 1983; Arno and Gruell 1986; see chapter 11.

4. Johnson 1950; Medin 1960; Miller 1964; Brotherson et al. 1984;

32. Fisher et al. 1987.

the two species of Cercocarpus occur together on the west side of


Flaming Gorge Reservoir.
5. Vlamis et al. 1964; Youngberg and Hu 1972; Lepper and Flesch
ner 1977.
6. Duncan 1975; Young and Bailey 1975; Gruell et al. 1985; Arno

33. Major and Rejmanek (1992) view the two species of serviceberry
as one, citing several taxonomic studies.
34. Brown 1971.
35. Ludwig 1969; Mueggler 1975; Despain 1990.
36. Whitman and Hanson 1939; Hansen and Hoffman 1988.

and Wilson 1986; Johnson (1950) and Miller (1964) report on

37. Sauer 1978.

sprouting and layering, but new stands also become established

38. Welsh 1957; Baker and Kennedy 1985; Knight et al. 1987.

from seed.

39. MacCracken et al. 1983; Uresk and Boldt 1986.

7. True mountain-mahogany is thought to sprout more readily


than curlleaf mountain-mahogany.

40. Neilson and Wullstein 1983.


41. Neilson and Wullstein 1983; Harper et al. 1985.

8. Billings 1954.

42. Bartlein et al. 1997.

9. Wight and Fisser 1968; Spaeth 1981; Waugh 1986.

43. Harper et al. 1985.

10. Soils were shallow in all of Wight and Fissers (1968) 55 study

44. Krebill 1972; Beetle 1974a,b; Gruell and Loope 1974; DeByle

areas, with 3087 percent sand and a pH ranging from 6.8 to

1979; Olmsted 1979; Boyce 1989; Coughenour and Singer 1991;

8.1. Soil conductivity was low (0.451.50 dS/m).

Hessl and Graumlich 2002.

11. Romme et al. 2009.

45. Elk damage mature trees through antler rubbing or eating bark,

12. Romme et al. 2009.

which facilitates the spread of fungal diseases, such as aspen

13. Waugh 1986; similarly, Soul et al. (2004) concluded that graz-

canker; Parker and Parker 1983; Veblen and Lorenz 1986.

ing and an increase in nurse shrubs facilitated western juniper

46. DeByle 1979.

expansion in parts of the Great Basin. For further discussion of

47. Veblen and Lorenz 1986.

sagebrush and livestock grazing, see chapter 7.

48. Coughenour 1991a.

14. Burkhardt and Tisdale 1976; Wright et al. 1979; Young and
Evans 1981.
15. Waugh 1986; Miller et al. 1993.
16. Romme et al. 2009.
17. Lyford et al. 2003; Romme et al. 2009.

49. Smith 2001, 2012.


50. Vander Wall and Balda 1977, 1983; Ligon 1978; Tranquillini
1979; Tomback 1983; Arno and Hammerly 1984.
51. Bamberg 1971; Hutchins and Lanner 1982; Arno and Hammerly
1984.

18. Baker and Shinneman 2004; Gray et al. 2006; Romme et al. 2009.
19. Gray et al. 2004, 2006; Romme et al. 2009; Breshears et al.
(2005) observed widespread pinyon pine mortality in the
southwest stemming from to global-change-type drought.
20. Knapp and Soul 2008.
21. Welsh 1957; Foster 1968; Everett 1987; Miller et al. 1993; Gray et
al. 2004.

Chapter 11: Mountain Forests


1. Daubenmire 1943b; Knight et al. 1977.
2. MacMahon and Anderson (1982) adopted a general adiabatic
lapse rate of 3F per 1,000 feet, but rates can vary with time of
year.

337

338

Notes to Pages 173188


3. Smith and Geller 1979; Smith and Knapp 1990.

32. Bentz et al. 2010.

4. Weaver and Perry (1978) found lodgepole pine on both sand-

33. Lotan and Perry 1983; Christiansen et al. 1987; Amman 1989;

stones and shales in the Bridger Range of south-central Montana.


5. Freeland 1944; Parker 1953; Perry 1971; Smith and Knapp 1990.
6. Kyte 1975.
7. Mooney and Billings 1960; Kimball et al. 1973; Salisbury 1984;
Hamerlynck 1992.

Bartos and Amman 1989.


34. The characteristics of high- and low-vigor lodgepole pine trees
are described by Kaufmann and Watkins (1990).
35. Waring and Pitman 1985.
36. Raffa et al. 2008; Bentz et al. 2010.

8. Sakai and Weiser 1973.

37. Bentz et al. 2010.

9. Knight 1991; Schoettle 1991.

38. The blue-stain fungus does not affect wood strength and is

10. Harvey et al. 1987.


11. Cui 1990.
12. Rochow 1970; Linhart et al. 1981; Knowles and Grant 1983;
Mitton 1985; Rehfeldt 1985; Schuster et al. 1989.

named for the bluish color it gives to the sapwood; some prefer
the more varied grain pattern of beetle wood.
39. Edburg et al. 2012.
40. Knight et al. 1991.

13. Buskirk et al. 1999.

41. Bentz et al. 2010.

14. Hibernation is a commonly used term, but there is ongoing

42. Regniere and Bentz 2007; Chapman et al. 2012. Preisler et al.

debate about which animals actually have extended slow-wave

(2012) concluded that warmer winter temperatures and drought

sleep and temperature-independent cellular depression over


extended periods of weeksthe definition of hibernation.
15. Meyer et al. (2005) and Dillon et al. (2005) synthesize informa-

were the strongest predictors of a beetle outbreak.


43. Amman 1989; Bartos and Amman 1989.
44. Raffa et al. 2008; Powell et al. 2012; Simard et al. 2012.

tion on the historical range of variability for the forests of the

45. Raffa et al. 2008; Bentz et al. 2010.

Bighorn and Medicine Bow mountains.

46. Klutsch et al. 2009; Kayes and Tinker 2012; Nelson et al. 2014.

16. In contrast, Severson and Kranz (1976) found that understory

The term stand is commonly used for a specific tract of forest.

plant growth in aspen forests was not correlated with tree den-

47. Collins et al. 2011; Diskin et al. 2011; Kayes and Tinker 2012.

sity or basal area.

48. Romme et al. 1986 (Yellowstone); Collins et al. 2011 (Colorado).

17. Kyte 1975; Knight et al. 1977.

49. Collins et al. 2011.

18. Hawksworth and Johnson 1989.

50. Pelz and Smith 2012.

19. In the first edition of this book, published in 1994, the mistle-

51. Klutsch et al. 2009; Diskin et al. 2011; Kayes and Tinker 2012.

toe was identified as the single most important problem in

52. Using Daubenmires approach, Wyoming forests have been

lodgepole pine forests. Today bark beetles have that distinc-

divided into 9 general categories (series) and 110 habitat types

tion. See also Hawksworth and Johnson 1989.

and community types, including 9 different kinds of ponderosa

20. Geils and Jacobi 1990; Jacobi et al. 1993.

pine forest, 16 kinds of lodgepole pine forest, and 28 kinds of

21. Smith 1970, 1975; Benkman et al. 1984; Siepielski and Benkman

subalpine forest. Considering such detail is beyond the scope of

2007, 2008.

this book. For descriptions of the Bighorn and Medicine Bow

22. Benkman et al. 2012; see also Smith 1970, 1975.

National Forests, see Meyer et al. (2005) and Dillon et al. (2005);

23. Parchman et al. 2012; Talluto and Benkman 2013.

see also the websites of other national forests. Reed (1971, 1976)

24. Langenheim 1962; Patten 1963; Despain 1973; Veblen and


Lorenz 1986; Green and Conner 1989; Peet 2000.
25. Minckley et al. 2012; Bartlein et al. (1997) discuss the potential

reports on the forests of the Wind River Mountains.


53. Cochran and Berntsen 1973.
54. Bartlein et al. 1997.

effects of climate change on tree species distribution in Yellow-

55. Baker 2009.

stone National Park.

56. Sherriff and Veblen 2006; Brown et al. 2008.

26. For example, Veblen and Lorenz (1986) describe human-caused


disturbances in relation to gold and silver mining in the Front
Range of Colorado.
27. Active suppression of fire did not become consistently effective

57. Baker et al. 2007; Sherriff and Veblen 2007; Schoennagel et al.
2011.
58. Shinneman and Baker 1997.
59. Ponderosa pine re-establishment was extremely slow on some

until the mid-1900s in most parts of the West, whereas grazing

steep, mid-elevation, south-facing slopes after a severe Colo-

and other land uses began to limit fire spread as early as the late

rado fire in 1851 that was followed by a relatively dry decade.

1800s. Therefore ecologists usually refer to fire exclusion during

Fire-killed snags were still conspicuous, and ponderosa pine

the past century to account for the effects of both suppression


and land use changes.
28. Heterogeneous fire severity was apparent, for example, in the
1988 Yellowstone fires (Turner et al. 1994) and in the 2002 Hayman fire in Colorado (Finney et al. 2003).
29. Hutto 2008; Saab et al. 2014.

trees were sparse 148 years after the fire (Brown et al. 1999).
60. Brown and Wu 2005; Brown 2006; League and Veblen 2006.
61. Nacify et al. 2010.
62. Patten 1963; Despain 1973; Loope and Gruell 1973; Lanier
1978; Weaver and Perry 1978; Veblen and Lorenz 1986; Goldblum and Veblen 1992; Wyomings Douglas-fir is a different

30. Taylor and Barmore 1980.

subspecies than the Douglas-fir of coastal California, Oregon,

31. Veblen et al. 1991b,c; Brunelle et al. 2008; Raffa et al. 2008.

and Washington. The inland variety does not grow as large as

Notes to Pages 189202


its northwestern cousin and is more tolerant of a drier, colder

95. Peet 2000.

climate.

96. McDonough 1985; Kay 1993.

63. Arno and Gruell 1986, Goldblum and Veblen 1992; Baker 2009.

97. Mock et al. (2013) found greater genetic diversity in aspen

64. Loope and Gruell 1973; Littell 2002; Korb 2005.

stands than was previously thought to exist, but the patches

65. Amman and Ryan 1991; Donato et al. 2013. The western spruce

of aspen that all leaf out in the spring or turn color at the same

budworm is a moth. In the Rocky Mountains, the larvae feed


primarily on the buds and young leaves of Douglas-fir. Outbreaks appear to occur when many of the trees are in a state of
low vigor.

time in the fall are usually dominated by a single genet.


98. Reed 1971; Youngblood and Mueggler 1981; Brown and DeByle
1989; Bartos et al. 1991; Peet 2000.
99. Shinneman et al. (2013) emphasize the great variability that

66. Korb 2005.

exists among aspen fire regimes in the western United States

67. Green and Connor 1989.

and propose five general types of aspen fire regimes in relation

68. Moir 1969; Despain 1973, 1983, 1990; Whipple and Dix 1979;

to environment and associated species.

Lotan and Perry 1983; Parker 1986; Peet 2000. Lodgepole pine

100.Seager et al. (2013) review the literature on effects of herbi

may be the only conifer species that can tolerate some infertile

vory on aspen forests and emphasize that the relationships are

sites.
69. Romme 1982.

complex, highly variable, and not well understood.


101. Kulakowski et al. (2013) review the wide range of settings in

70. Johnson and Gutsell 1993.

which aspen grows and the equally wide variety of relation-

71. Lotan 1964. Logging slash is sometimes scattered across the

ships with conifers and other woody species. Many early stud-

ground after a lodgepole pine clearcut; the serotinous cones in

ies of aspen were conducted in specific stands over periods of a

the slash open after a few weeks and release seed, which re

few decades, often in places where aspen was replaced by other

generates the stand (Aoki et al. 2011).

species. Recent studies over larger areas often reveal strong

72. Lotan 1975.

aspen persistence. Wiens et al. (2012) review potential pitfalls

73. Tinker et al. 1994.

of using historical conditions as benchmarks for evaluating cur-

74. Muir and Lotan 1985a,b; Schoennagel et al. 2003.

rent ecological conditions.

75. Benkman and Siepielski 2004; Benkman et al. 2012; see note 23.

102. Manier and Laven 2002; Kulakowski et al. 2004, 2013.

76. Muir and Lotan 1985a.

103. Houston 1973; Loope and Gruell 1973; Beetle 1974b; Walters

77. Turner et al. 1997, 2004; Schoennagel et al. 2003.

et al. 1982; DeByle et al. 1987; Boyce 1989; Seager et al. 2013.

78. Kashian et al. 2005.

104. Worrall et al. 2010; Anderegg et al. 2013.

79. Kipfmueller and Baker 2000; Schoennagel et al. 2004.

105.Kashian et al. 2007; but see Worrall et al. (2010) and Forest

80. Westerling et al. 2006, 2011; but see Gavin et al. 2007.
81. Running 1980; Knapp and Smith 1981, 1982; Kaufmann 1985;

Ecology and Management (2013, vol. 299, pp. 1100).


106. Saab et al. 2014.

Smith and Knapp 1990. As discussed in chapter 3, seedlings are


more sensitive to environmental conditions than mature plants
are. In addition to the widespread Engelmann spruce and sub

Chapter 12: The Forest Ecosystem

alpine fir, white spruce is common in the Black Hills, and white

1. Bormann 2000 (pp. 45); see also Chapin et al. 2012.

fir is known from two localities in southwestern Wyoming (in

2. Net primary productivity often is reported for aboveground

the foothills of the Uinta Mountains near Lonetree and on

plant growth only because of the difficulty of estimating the

Little Mountain in Sweetwater County).

growth of roots. Pearson et al. (1987) estimated that total net

82. Sibold et al. 2006.

primary productivity in mature lodgepole pine forests of the

83. Romme and Knight 1981; Johnson and Fryer 1989; Veblen et al.

Medicine Bow Mountains, including roots, is only 2.53.2

1991a.

metric tons/hectare/year. Forests dominated by spruce and fir

84. Bartlein et al. 1997; Minckley et al. 2012.

probably are similar. Somewhat higher productivity would be

85. Schimpf et al. 1980; Pearson et al. 1987.

expected in younger forests (see fig. 12.10); forests on more fer-

86. Aplet et al. 1988.

tile soils, such as on floodplains; or those at lower elevations,

87. Schmid and Hinds 1974; Alexander 1987a; Baker and Veblen

where the growing season is longer. Site index is an indirect

1990; Veblen et al. 1991b,c.

measure of the productive potential of an area and is expressed

88. Ciesla 2011.

as the height of dominant trees at a specified age, typically 50

89. Oosting and Reed 1952; Shea 1985; Veblen 1986a,b; Alexander

or 100 years old. Sites having a higher site index are generally

1987a; Veblen et al. 1991a; Peet 2000.


90. Oosting and Reed 1952; Loope and Gruell 1973; Alexander
1987a; Cui 1990.
91. Veblen 1986a,b.

more productive than those with a lower site index.


3. Fahey (1983) estimated that the amount of decomposition
under snow is about equal to that during the rest of the year;
Brooks and Williams 1999.

92. Kaufmann 1985.

4. Fogel and Trappe 1978; Maser et al. 1978, 1988; McIntire 1984.

93. Aschan et al. 2001.

5. Vogt et al. 1982; many mycorrhizal fungi produce mushrooms,

94. Severson 1963.

the spore-producing part of a fungus.

339

340

Notes to Pages 202214


6. Allen 1991; Miller and Allen 1992.

costs associated with accumulation of wood and other non-

7. Harder 1979.

photosynthetic tissues reduced the net addition of new bio-

8. Knight et al. 1985.

mass. However, recent experiments by Ryan et al. (2004) do

9. For this unit of measure, the volume of water depends on the size

not support this explanation. Instead, multiple mechanisms

of the area being considered. Six inches of water, for example, is

may be involved, including reduced gross primary productiv-

the volume of water that will cover a specified area 6 inches deep.

ity, reduced canopy leaf area, and reduced growth efficiency

10. Knight et al. 1985; Fahey et al. 1988.


11. In lodgepole pine forests of the Medicine Bow Mountains,

(amount of wood produced per unit of resource used).


32. Smithwick et al. 2009a,b.

Knight et al. (1985) found that transpiration commonly

33. Knight et al. 1991.

accounts for 5061 percent of total annual evapotranspiration,

34. Minshall et al. 1997; Rhoades et al. 2011.

944 percent of which occurs during the spring drainage period

35. Chapin et al. 2002.

while snow still covers the ground (vernal transpiration). Esti-

36. Minshall et al. (1997) report that nitrogen levels in streams

mated vernal transpiration and outflow varied considerably

draining steep terrain underlain by friable volcanic substrates

among the stands that were studied, with vernal transpiration

were initially elevated after the 1988 Yellowstone fires, but post-

accounting for 420 percent of the snow water.

fire levels in other streams draining gentler topography were

12. Grier and Running 1977; Binkley et al. 1995.

barely detectable; see also Turner et al. 2007.

13. Knight et al. 1985.

37. Smithwick et al. 2009a,b.

14. Knight et al. 1985.

38. Pearson et al. 1987; Knight et al. 1991.

15. Kaufmann 1985; Kaufmann et al. 1987; Pataki et al. 2000.

39. A study in the Greater Yellowstone Ecosystem (Romme et al.

16. Jaynes 1978; Knight et al. 1985.

1986) found that annual wood production per unit area returned

17. Gary and Troendle 1982; Troendle 1983, 1987; Knight et al. 1985;

to previous levels or was even higher 1015 years after a bark

Troendle and Kaufmann 1987; but see Biederman et al. (2012)

beetle outbreak. Also, Brown et al. (2010) concluded that some

and the subsequent discussion of bark beetles in this chapter.

stands in British Columbia were still carbon sinks even during

18. MacDonald and Stednick 2003.

the first several years after severe bark beetlecaused mortality;

19. Eighty percent of the roots are in the top 8 inches of the soil.

and Pfeifer et al. (2011) projected that the amount of ecosystem

20. Gary and Troendle 1982.

carbon in Idaho forests would recover in 25 years or less after a

21. Nsholm et al. 1998.

bark beetle outbreak. Litton et al. (2003) studied the importance

22. Fahey et al. 1985; Knight et al. 1985; Fahey and Knight 1986.

of soil microbes and organic matter in the carbon budget.

23. Gorham et al. 1979.

40. Edburg et al. 2012; Pugh and Gordon 2012.

24. Harmon et al. 1986.

41. Pugh and Small 2012. Clow et al. (2011) observed higher soil

25. Gosz 1980; Fahey 1983; Prescott et al. 1989; Knight 1991.

moisture beneath beetle-killed trees in northern Colorado,

26. Fahey 1983.

which they interpreted as reduced evapotranspiration; similar

27. Some changes caused by clearcutting are similar to those caused

changes were documented in lodgepole pine forests of British

by fire. Stottlemyer and Troendle (1999) studied the effects of

Columbia by Boon (2012).

clearcutting on snowpack, soil solution chemistry, and nutrient

42. Biederman et al. 2012; Pugh and Gordon 2012.

export on the Fraser Experimental Forest in Colorado.

43. Edburg et al. 2012.

28. The proportion of the surface soil that is sterilized by the heat
of fires in the Rocky Mountain region is small and is soon mitigated by natural processes that facilitate occupancy by new
organisms.

44. Morehouse et al. 2008; Clow et al. 2011; Griffin et al. 2011; Griffin and Turner 2012.
45. Mikkelson et al. (2013a,b) found that mountain pine beetles
caused some changes in streamwater chemistry.

29. Turner et al. (2004) found that, in young forests 10 years after

46. Ecosystems become a source of carbon dioxide when the carbon

the 1988 Yellowstone fires, approximately 68 percent of the

fixed by photosynthesis is less than the carbon released by the

burned landscape was characterized by an aboveground net

respiration of plants and all other organisms.

primary production of less than 2.0 tons/hectare/year; 22 per-

47. Edburg et al. 2012.

cent by 24 tons/hectare/year; and 10 percent by greater than

48. Troendle 1987; Troendle and Nilles 1987; Prescott et al. 1989a;

4 tons/hectare/year. The tree stratum accounted for most of the


leaf area and productivity in almost all stands, even in those
with low tree densities.

Knight et al. 1991.


49. Intense fires in some areas can lead to a water-repellent surface,
which reduces infiltration and increases surface runoff and

30. Kashian et al. 2005. In the Medicine Bow Mountains, Pearson

possibly erosion. In burned Rocky Mountain forests, water-

et al. (1987) estimated that maximum total net primary produc-

repellent soils tend to be patchy, cover a small area, and remain

tivity rates of 2.53.2 tons/hectare/year are reached in stand


ages of 4060 years.
31. It was long thought that the decline in net primary productiv-

repellent only for several months at most.


50. Harmon et al. 1986; Maser et al. 1988; Tinker and Knight 2004;
Krzyszowska-Waitkus et al. 2006.

ity as forests age was because, although photosynthesis contin-

51. Tinker and Knight 2004; Krzyszowska-Waitkus et al. 2006

ued to produce sugars at previous levels, increasing respiration

52. Tinker and Knight 2001, 2004.

Notes to Pages 214228


53. Romme and Knight 1982; Minshall et al. 1989.

79. Dillon et al. 2005; Meyer 2005.

54. Alexander 1986, 1987a,b.

80. Fettig et al. 2007.

55. Harmon et al. 1986; Bull et al. 1997.

81. Raffa et al. 2008.

56. There is growing interest in the idea of emulating natural distur-

82. Collins et al. 2012.

bance patterns and processes in silvicultural operations (Kohm

83. Worrall et al. 2010.

and Franklin 1997; Perera et al. 2004; Dillon et al. 2005; Meyer

84. Anderegg et al. 2013.

et al. 2005). Because the ecosystems are adapted to these kinds

85. Romme et al. 2011.

of disturbances, they are likely to respond well to anthropo

86. Zier and Baker 2006.

genic disturbances that mimic the natural disturbances. Other

87. See http://forest.moscowfsl.wsu.edu/climate/species/index.php

kinds of anthropogenic disturbances that have no natural prec-

for a summary of research done by Gerald Rehfeldt and his

edent are more likely to result in undesirable responses.

associates; see also Bartlein et al. 1997.

57. Knight et al. 1991; Parsons et al. 1994; killing 60 percent of the

88. Millar et al. (2007) describe approaches for adapting to climate

trees by girdling reduced leaf area by 43 percent and led to an

change, such as by conserving ecosystems that are resilient to

estimated 92 percent increase in water outflow compared to

climate impacts or assisting the migration of threatened species

clearcutting, which increased water outflow by 277 percent; see

to new habitats. They also suggest ways to mitigate climate

also Biederman et al. 2012; Rhoades et al. 2013.

change, such as by increasing carbon storage in forests, to the

58. Coops and Waring 2011.

extent possible.

59. Romme and Turner 1991.


60. Rehfeldt et al. 2006; see also chapter 16.
61. MacDonald and Stednick 2003.
62. Westerling et al. (2006) and Wotton et al. (2010) write about
fire; see Raffa et al. (2008) for more on beetle outbreaks.
63. Simard et al. 2012; Mitton and Ferrenberg (2012) found that
mountain pine beetles in Colorado now produce two generations per year rather than one, due to climate warming. For a
specific example and a review of the literature, see Raffa et al.
2008.
64. Powell et al. 2012. A somewhat different conclusion might be

Chapter 13: Mountain Meadows and Snowglades


1. Debinski et al. 2000.
2. Ellison 1954; Starr 1974; Windell et al. 1986; Debinski et al.
2000.
3. Dunnewald 1930; Daubenmire 1943a; Jackson 1957; Hurd 1961;
Patten 1963; Despain 1973; Dunwiddie 1977; Peet 2000.
4. Dunnewald 1930; Behan 1957; Patten 1963; Cary 1966; Miles
and Singleton 1975.
5. Miles and Singleton 1975; Vale 1978.

reached for ponderosa pine or Douglas-fir forests, where greater

6. W. K. Smith et al. 2003; see also Sibul (1995) on Dry Park.

numbers of moderately injured trees often remain after the

7. Munroe (2012) found no clear differences in the soil par-

fires of variable severity that typically characterize these forest

ent material under the meadow and forest. He classified the

types.

meadow soils as Typic Humicryepts and the forest soils as

65. Bentz et al. 2010; little is known about the fir bark beetle.

Inceptic Haplocryalfs, and assumed the differences could have

66. Hicke et al. 2012; see also Klutsch et al. 2011; chapter 11.

developed since the forest disturbance that created the park.

67. Compare Simard et al. 2011 and Schoennagel et al. 2012.

See also Coop and Givnish 2008.

68. See Simard et al. 2011; Schoennagel et al. 2012.

8. Buckner 1977.

69. Lynch et al. 2006; the spatial pattern of burning in the 1988

9. Httenschwiler and Smith 1999; Moir et al. 1999.

Yellowstone fires was significantly correlated with the spatial pattern of a previous mountain pine beetle outbreak that occurred

10. Bekker and Malanson (2008) review the types and causes of ribbon forests.

during 19721975, with the odds of burning increasing by 11

11. Billings 1969 (p. 204).

percent in areas of previous beetle activity compared with areas

12. Arno and Hammerly 1984; Httenschwiler and Smith 1999;

without beetles. However, no correlation was found between


burning and more recent beetle activity during 19801983.

Bekker and Malanson 2008; Bekker et al. 2009.


13. Buckner 1977; Bekker et al. 2009.

70. Kulakowski and Jarvis 2011; Black et al. 2013.

14. Bekker and Malanson 2008.

71. Westerling et al. 2006; Raffa et al. 2008; Logan et al. 2010.

15. Coughenour 1991a,b; Coughenour and Singer 1991.

72. Wallace 2004; Romme et al. 2011.

16. Stohlgren et al. (1999) studied grazing exclosures in the Rocky

73. Millspaugh et al. 2004.

Mountain region, finding that plant species richness was just as

74. Coops and Waring 2011.

great inside exclosures as outside and that grazing in their study

75. Firewise guidelines are available from many sources, includ-

area had no or little effect on the presence of exotic species.

ing http://www.firewise.org/resources/files/Around-Your-Home

17. Billings 1969; Moir et al. 1999; W. K. Smith et al. 2003.

.pdf.

18. Lynch (1998) studied Fish Creek Park in the Wind River Moun-

76. Cohen and Stratton 2003; for a booklet titled Living with Wildfire in Wyoming, see www.uwyo.edu/barnbackyard.

tains; see chapter 2.


19. Jakubos and Romme (1993) found no charred stumps or other

77. Schoennagel and Nelson 2011.

wood in the meadows they studied; Munroe (2003) worked in

78. Finney 2001.

the Uinta Mountains.

341

342

Notes to Pages 228240


20. Anderson and Baker (2005) could not rule out the influence of

(Indian paintbrush). Such plants are known as alternate hosts;

livestock grazing in the Medicine Bow Mountains, as Jakubos

see Zambino et al. 2005; McDonald et al. 2006; http://www

and Romme (1993) could in Yellowstone National Park.


21. Debinski et al. 2010.
22. Forrest et al. 2010.

.fs.fed.us/rm/highelevationwhitepines/Threats.
17. Resler and Tomback 2008; Keane and Schoettle 2011; Larson
2011.

23. Harte and Shaw 1995; Shaw et al. 2000.

18. Billings 1973; Pw 1983.

24. Price and Waser 2000.

19. Smith and Geller 1979; Tranquillini 1979; Hadley and Smith

25. Price and Waser 2000.


26. Debinski et al. 2000.

1986.
20. Scott 1995.
21. Exceptions are the small annual plants studied by Reynolds

Chapter 14: Upper Treeline and Alpine Tundra


1. Peet 2000; Krner 2012; Daubenmire suggested in 1954 that
alpine treeline drops approximately 350 feet per degree of latitude between 30 and 60 north latitude.
2. Daubenmire 1954; west slopes tend to have more snow, because

(1984) in the Beartooth Mountains: Koenigia islandica, Poly


gonum confertifolium, and P. douglasii. Also, see Forbis (2003),
who found abundant seedlings of some plants in the alpine
tundra on Niwot Ridge in Colorado.
22. Terjung et al. 1969; Tieszen and Detling 1983; Bowman and Fisk
2001.

most storms come from the west, with precipitation occurring

23. Scott and Billings 1964; Sakai and Otsuka 1970; Kimball et al.

as the air rises on the western slopes. Generally, annual precipi-

1973; Wallace and Harrison 1978; Bell and Bliss 1979; Hamer-

tation is less on the leeward, east-facing slopes, stemming from


the rainshadow effect (see chapter 3).
3. Whitebark pine is a close relative to limber pine, with both having five needles per fascicle and many of the same ecological
characteristics; Driese et al. (1997) calculated a mean elevation
of 9,725 feet for upper treeline in Wyoming.

lynck 1992.
24. Krner 2012.
25. Rochow 1970.
26. Spomer 1964; Campbell 1997 (plants); Molenda et al. 2012
(insects).
27. Knutson 1981.

4. W. K. Smith et al. 2003.

28. Armstrong et al. 2001; Dearing 2001.

5. Krner (2012) describes factors that limit growth at treeline,

29. Huntly 1987.

whereas W. K. Smith et al. (2003, 2009) and Wieser et al. (2014)

30. Huntly and Inouye 1988; Armstrong et al. 2001; Seastedt 2001.

focus on factors limiting seedling establishment.

31. On Niwot Ridge in Colorado, the plant communities are classi-

6. Layering is the process whereby a new tree develops when low

fied as fellfield, dry meadow, moist meadow, wet meadow, late-

branches of spruce and fir produce roots while pressed to the soil

melting snowbank, and shrub tundra (Walker et al. 2001); see

surface during the winter by heavy snow. This is an example of


vegetative reproduction that results in a clone. See fig. 11.18.
7. Marr 1977; Benedict 1984; Holtmeier and Broll 1992; Seastedt

Thilenius and Smith 1985; Moir et al. 1999.


32. Billings (2000) suggested three terms that now are widely used
when describing vegetation patterns in the alpine tundra:

2001 (on soils); Seastedt and Adams 2001.

macrog radients, microgradients, and mesogradients. See also

8. Seastedt et al. 2004; Hiemstra et al. 2006.

Johnson and Billings 1962; Oberbauer and Billings 1981; Litaor

9. Griggs 1946; Daubenmire 1954; Wardle 1968; Tranquillini


1979; Arno and Hammerly 1984; Krner 2012.

et al. 2002, 2008.


33. Though freeze-thaw cycles still affect alpine tundra, some

10. Krner 2012; previously treeline was thought to occur where

patterned ground may be remnants of the Pleistocene, as sug-

the July mean temperature is less than 50F (Tranquillini 1979;

gested by solifluction terraces found in mountain foothills

Arno and Hammerly 1984).

and soil polygons in some intermountain basins (see chapter

11. Smith and Carter 1988.


12. Bristlecone pine also has flexible branches and is found in Colorado, northern New Mexico, and westward.

3). See Richmond (1949) for an early study in the Wind River
Mountains.
34. Billings and Mooney (1959) studied cryoturbation in the Medi-

13. Hadley and Smith 1986, 1987, 1989.

cine Bow Mountains, where in a few square yards the tundra

14. Eversman 1968; Potter 1969; Mears 1975; Malanson and Butler

changed from a frost hummock to a sorted polygon and back

1984; Butler 1985; Johnson 1987; Patten and Knight 1994.


15. Hoff 1992; Tomback et al. 2001; Keane and Schoettle 2011; white

to another frost hummock. Phillips (1982) observed similar


patterns associated with stone polygons in Rocky Mountain

pine blister rust causes swelling on the branches and trunk, even-

National Park.

tually killing the tree. In general, the disease infects five-needle

35. Willard et al. 2007.

pines, known collectively as white pines; see chapter 15. Bristle-

36. Bowman et al. (2006) concluded that changes in nitrogen

cone pine is found in central Colorado but not in Wyoming.


16. Pathogenic fungi known as rusts often require two unrelated

uptake after fertilization occur partially because of shifts in


species composition.

species of host plants to complete their life cycle. For white

37. Bowman and Fisk 2001; Monson et al. 2001; as in forests, some

pine blister rust, the fungus requires plants in the genus Ribes

nitrogen is absorbed by the roots in organic forms, such as the

(wild currant or gooseberry), Pedicularis (lousewort), or Castilleja

amino acid known as glycine.

Notes to Pages 240257


38. Baron et al. 1994; Baron and Campbell 1997; Stottlemeyer et al.

15. Strong 1968 (p. 51).

1997; Heuer et al. 1999; Burns 2003, 2004.

16. Janetski 1987 (chapters 2 and 3); Frison 1991; Kornfeld et al.

39. Fenn et al. 2003; Porter and Johnson 2007.

2010; artifacts found on the northern end of Jackson Lake doc-

40. Initially the primary source for nitrogen was thought to be the

ument hunting camps dating back about 2,500 years.

Denver metropolitan area and large agricultural operations to

17. Schreier 1982.

the east, including feedlots. However, further research sug-

18. Haines 1965.

gested that much of the atmospheric nitrogen is in the form of

19. Despain 1990.

nitrous oxides coming from the Salt Lake City area, as reported

20. Despain 1983, 1990.

by Sievering (2001).

21. Despain 1990.

41. Fenn et al. 2003; Hood et al. 2003; Seastedt et al. 2004; Liptzin
and Seastedt 2010.
42. Fenn et al. 2003.

22. Chadde et al. 1988.


23. Patten 1959; Despain 1990.
24. Mattson 1984; Lemly 2007.

43. Timothy Seastedt, pers. comm.

25. Chadde et al. 1988; Chadde and Kay 1991.

44. Carrera et al. (1991), who calculated that the upper treeline

26. Sheppard 1971.

could rise 260460 feet with less than a 2F increase in the

27. Fryxell 1978.

mean summer temperature; Lynch 1998; Elias 2001; Toney and

28. Righter 1982.

Anderson 2006.

29. Palmer 1991.

45. Billings (1969) found evidence of fire at treeline in the Medicine


Bow Mountains, which had the effect of lowering the treeline,
because new tree establishment was slow.
46. W. K. Smith et al. 2003, 2009.
47. Kupfer and Cairns 1996; Malanson et al. 2007; Loffler et al.
2011.

30. Schreier 1982; Righter 1982.


31. Greater Yellowstone Coordinating Committee report of 1987,
available at http://fedgycc.org/index.html; Youngblood and
Mueggler 1981.
32. Steele et al. 1983.
33. Oswald 1966; Love et al. 2003.

48. Weisberg and Baker 1995a,b.

34. Loope and Gruell 1973.

49. Baker and Weisburg 1995, 1997.

35. Sabinske and Knight 1978.

50. W. K. Smith et al. 2003.

36. Blackwelder 1912; Voight 1982; Lawrence and Lawrence 1984.

51. Forbis 2003; Elliott and Kipfmueller 2010, 2011.

37. Pritchard 1999; Schullery 2004.

52. Munroe 2003; Stinson 2005; Cannone et al. (2007) worked in

38. Schullery 2010. Comstock wrote about the value of wildlands

the Alps, finding rapid upward treeline advance.

for evolution in the 1870s, at a time when such ideas were new

53. Inouye 2008; Wipf et al. 2009; Loffler et al. 2011; in Alaska,

and controversial. Wright and Leopold championed predators

yellow cedar is dying in some places because shallow roots are

in the 1930s. George Bird Grinnell described Yellowstone in the

killed by frost, due to reduced insulation from snow (Hennon et

late 1800s and early 1900s as a reservoir of water and wildlife

al. 2012).

which could sustain surrounding lands.


39. Leopold et al. 1963.

Chapter 15: The Greater Yellowstone Ecosystem

40. Schullery 2004.


41. The decade of the 1870s probably was the period of greatest

1. Eversman and Carr 1992; Hansen et al. 2002. Also see the Atlas

slaughter in and around YNP (Schullery 2004). The first super-

of Yellowstone (Marcus 2012) for a collection of maps pertaining

intendent of YNP, Nathaniel Langford, had almost no budget

to geography, wildlife, and ecology.

or staff and visited the park only twice during his five years

2. Schreier 1983; Good and Pierce 2002; Mogk et al. (2012) provide
a geologic map for YNP.

as superintendent; meanwhile, tourists and local residents


exploited park resources essentially without constraint. In 1886

3. Dorf 1964.

a troop of U.S. Cavalry was sent to protect YNP; the Cavalry

4. Good and Pierce 2002.

managed the park until creation of the National Park Service in

5. Love et al. 2003.

1916 (Schullery 2004).

6. Love et al. 2003.

42. Houston 1982.

7. Good and Pierce 2002.

43. Pritchard 1999; Schullery 2004.

8. Good and Pierce 2002; Love et al. 2003.

44. Kay 1993.

9. Whitlock 1993; Love et al. 2003; Millspaugh et al. 2004.

45. Ecologists Andrew Hansen and Jay Rotella (2002) found that the

10. Marston and Anderson 1991.

greatest diversity and abundance of birds in the GYE are associ-

11. Haines 1996; Schullery 2004.

ated with aspen, cottonwood, and willow habitats. In another

12. Helena Daily Herald, November 9, 1870 (p. 1).

study, Berger et al. (2001) compared bird communities in Grand

13.
Haines 1996; http://www.cr.nps.gov/history/online_books/

Teton National Park, where moose were relatively abundant

haines1/iee2d.htm.
14. Meagher and Houston (1998) compared modern photographs
in YNP with those taken by early explorers.

and fed heavily on willows, to nearby national forests, where


moose numbers and browsing intensity were lower. Species
richness and abundance of nests were greater in the national

343

344

Notes to Pages 257264


forests. Also, two riparian specialists, the gray catbird and Mac
Gillivrays warbler, were absent in the park, where moose densities were high, but were present in the national forests.

61. Yellowstone National Park 1997; National Academy of Sciences


2002.
62. Schullery 2004.

46. Schullery and Whittlesey 1992.

63. Despain 1990; Renkin and Despain 1992.

47. D. W. Smith et al. 2003.

64. Schullery 1989.

48. Smith 2001, D. W. Smith et al. 2003; Kilpatrick et al. 2009;

65. Christensen et al. 1989.

Cross et al. 2010; Bienen and Tabor 2006; brucellosis and other

66. Romme and Despain 1989.

diseases have not been a problem for elk and bison.

67. Millspaugh et al. 2000, 2004; Higuera et al. 2011.

49. Chase (1986), Bonnicksen (1989), and Wagner (2006) were

68. Shullery 1989; several other fires also were ignited by humans,

major critics of park policy; Pritchard (1999) reviews some of

or were outside YNP boundaries, or both, and contributed to

the controversy.
50. Elk are not the only wildlife species adversely affected by land

the total acreage burned in the GYE in 1988.


69. Brown 1991.

use changes outside park boundaries. Hansen and Rotella (2002)

70. Wallace 2004; Romme et al. 2011.

analyzed nesting success of yellow warblers, neotropical migrants

71. Kay 1993; Romme et al. 1997, 2005; Stevens et al. 1999.

that nest in deciduous woodlands of aspen and cottonwoods,

72. Schullery 2004.

both inside and outside YNP. Woodlands in the park commonly

73. Westerling et al. 2006.

contained nesting yellow warblers, yet those birds often failed

74. Westerling et al. 2011.

to produce adequate numbers of surviving offspring to permit

75. For example, species that are now rare or absent in the GYE,

population growth because of low summer temperatures at the

such as ponderosa pine, could become more common, as noted

relatively high elevations that characterize park habitats. They


concluded that the park by itself cannot sustain a population of

by Bartlein et al. (1997) and Westerling et al. (2011).


76. Tomback et al. 2001.

the warblers; the birds that are found in the park have come from

77. Logan et al. 2010; Bockino and Tinker 2012.

more productive habitats elsewhere. Yellow warblers nesting in

78. Tomback and Achuff 2010; the rust kills other five-needle pines

woodlands at lower elevations, on private lands outside the park,


could potentially produce positive numbers of offspring under

also, such as limber pine (see chapter 14).


79. Bockino and Tinker 2012.

the warmer summer conditions in these habitats. However,

80. Logan et al. 2010; Tomback and Achuff 2010.

many woodlands on private lands are fragmented by agriculture

81. Keane and Schoettle 2011.

and exurban development, and this landscape pattern facilitates

82. Logan and Powell 2001.

nest predation by brown-headed cowbirds. As a result, yellow

83. http://www.greateryellowstonescience.org/download_product/

warblers generally fail to produce sufficient surviving offspring

538/0.

to sustain the population in these habitats as well. Apparently

84. Gresswell and Liss 1995.

the yellow warbler populations throughout the areas studied by

85. In 2012 more than 300,000 lake trout were removed from

Hansen and Rotella (2002) are sustained by birds coming into

Yellowstone Lake, and the number of juvenile cutthroat trout

this area from more suitable habitats elsewhere rather than by

appeared to be higher than in recent years (Yellowstone Associa-

local reproduction. See also Hansen and DeFries 2007.

tion E-Newsletter, December 2012).

51. National Academy of Sciences 2002.

86. Middleton et al. 2013d; see also Wilmers and Post 2006.

52. National Academy of Sciences 2002.

87. Schwartz et al. 2006.

53. Frank and McNaughton 1992, 1993. Deposition of urine and

88. Fortin et al. 2013; Middleton et al. 2013d.

feces, as well as soil scarification by the action of hooves, also

89. Mattson et al. 2002.

may contribute to high plant growth in grazed grasslands.

90. National Academy of Sciences 2002; Schullery 2004.

54. Coughenour 1991a,b; Coughenour and Singer 1991.

91. Pritchard 1999.

55. Houston 1982 (p. 198).

92. National Academy of Sciences 2002.

56. Chadde and Kay 1991.

93. Vucetich et al. 2005; seventeen additional Canadian wolves

57. Romme et al. 1995; Ripple and Larsen 2000.

were released in Yellowstone in April 1996.

58. Patten 1968; Singer et al. 1994; Yellowstone National Park 1997.

94. National Academy of Sciences 2002.

59. National Academy of Sciences 2002.

95. Middleton et al. 2013b,c,d; some evidence indicates that the

60. Research suggests that aspen persistence is not threatened in


the GYE as a whole. Brown et al. (2006) determined aspen cover

wolf population on the northern range of YNP has declined as


the number of elk has dropped.

across the GYE in 1956 and 2001; they found an average loss of

96. Vucetich et al. 2005.

10 percent of aspen cover overallfar less than was suggested

97. Beschta 2003, 2005; Ripple and Beschta 2004; Beschta and

by fine-scale studies of aspen that commonly have been conducted where ungulate browsing is conspicuous. Brown et al.
(2006) also determined that aspen is rare in the GYE, occupying only 1.4 percent of the region, and is more common in the
southern than in the northern part of the GYE.

Ripple 2010.
98. Beyer et al. 2007; Creel and Christianson 2009; Kauffman et al.
2010; Middleton et al. 2013a.
99. Kauffman et al. 2010; Middleton et al. 2013a. Kimble et al.
(2011) did find places where the density and size of aspen

Notes to Pages 265275


sprouts apparently had increased between 1991 and 2006, but

part of the Limestone Plateau. Tertiary igneous intrusions in

cautioned that these patterns cannot be extrapolated to the

the Hills were formed 55 to 33 million years ago and include

landscape or attributed solely to the effects of wolves. Eisenberg

Bear Butte, Black Butte, Crow Peak, Custer Peak, Devils Tower,

et al. (2013) review the literature on wolf-elk-aspen research

Inyan Kara Mountain, the Little Missouri Buttes, Sundance

and emphasize that general patterns and effects are not clear.
See also http://nyti.ms/1cHqikq.

Mountain, Terry Peak, and Warren Peaks (Lisenbee et al. 1981).


14. Shepperd and Battaglia (2002) provide a review of all forest

100.Marshall et al. 2013; where willows have recovered, Baril et

types; see also Larson and Johnson 1999; Marriott et al. 1999;

al. (2011) found an increase in some species of songbirds; Baker

Marriott and Faber-Langendoen 2000; Wienk et al. 2004;

et al. (2005) describe the effect of elk herbivory on willows and


beaver; Hebblewhite et al. (2005) describe how human activity
has mediated the trophic cascade caused by wolves.
101. Berger et al. 2008.
102. Hansen and Rotella 2002; Parmenter et al. 2003; Gude et al.
2007; Hansen and DeFries 2007; Bilyeu et al. 2008.
103.Olliff et al. 2010.
104. Schullery 2010 (p. 12).

Battaglia et al. 2008.


15. Progulske 1974; Bock and Bock 1984; see chapter 10 and the
condition reports produced by the various units of the National
Park Service (available online).
16. Boldt and Van Deusen 1974.
17. Marriott and Faber-Langendoen (2000) provide a more detailed
classification.
18. Marriott and Faber-Langendoen (2000) describe the riparian
and wetland plant communities of the Black Hills.

Chapter 16: The Black Hills, Bear Lodge Mountains, and


Devils Tower
1. The Lakota name for the Black Hills is Paha Sapa, which trans-

19. McIntosh 1931.


20. Severson and Thilenius 1976.
21. MacCracken et al. 1983; Marriott and Faber-Langendoen 2000;
see chapter 5.

lates to hills that are black, an impression created by the

22. The three characteristic grasses are Sporobolus heterolepis, Achna

appearance of ponderosa pine forests from a distance. Lageson

therum richardsonii, and Danthonia intermedia, respectively. Mar-

and Spearing (1988) and Gries (1996) summarize the geologic

riott (2012) describes the Black Hills montane grassland type,

history of the area and include road logs that identify geo-

which is considered unique to the Black Hills; http://www

logic features; Froiland (1990) and Raventon (1994) review the

.natureserve.org/explorer/servlet/NatureServe; the large grass-

natural history of the Black Hills. The geologic origin of Devils

lands in the Central Area near Deerfield Reservoir, known as

Tower is unclear, but it is one of several igneous intrusions in

Reynolds, Slate, and Gillette prairies, have been largely con-

the area (see endnote 13).

verted to hay meadows.

2. Froiland 1990. The Vrendrye brothers encountered several

23.
MacCracken et al. 1983; Hoffman and Alexander 1987;

tribes along the way, traveling most likely to the Bighorn Moun-

mountain-mahogany is found only in the Black Hills east of

tains on a route north of the Black Hills (http://en.wikipedia

Newcastle and is rare in the Bear Lodge Mountains.

.org /w i k i/ Ve re nd r ye _ Brot he r s _ jou r ney_ to _ t he _ Ro c k y_

24. Boldt and Van Deusen 1974; Shepperd and Battaglia 2002.

Mountains). The Sioux arrived later in the 1700s. A treaty was

25. Shepperd and Battaglia 2002; Spiering and Knight 2005; see

signed in 1868 that gave the Sioux rights to the Black Hills. However, ownership is still the subject of a legal dispute between the
Sioux and federal government (http://www.defendblackhills
.org).
3. Custer 1875 (p. 506); Donaldson 1875 (p. 564); Ludlow 1875;
Raventon 1994; Grafe and Horsted 2002.

also chapter 12.


26. Wienk et al. (2004) found that a combination of thinning and
prescribed burning stimulates understory plant growth.
27. Uresk and Paintner (1985) studied cattle and found that shrubs
and trees in the northern Black Hills provide 37 percent of
their diet during September; in summer the diet was 54 per-

4. H. Marriott, pers. comm.

cent grasses, 17 percent forbs, and 28 percent shrubs and trees.

5. Progulske 1974; Grafe and Horsted 2002.

Elsewhere, cattle consume recently fallen cottonwood leaves in

6. McIntosh 1931; Raventon 1994; Larson and Johnson 1999.

September and October, such as along the South Platte River in

7. Buttrick 1914; Wright 1970.

Colorado (Sedgwick and Knopf 1991).

8. King et al. 2013.


9. Larson and Johnson (1999) describe six geomorphic regions

28. Fisher et al. 1987; Wienk et al. 2004; Brown 2006; Brown et al.
2008.

rather than five, adding the Gray Shale Foothills common

29. Shinneman and Baker 1997.

around the Bear Lodge Mountains.

30. Newton and Jenney 1880 (p. 322).

10. The Spearfish Formation is known as the Chugwater Formation


in Wyoming; gypsum also is found in the formation.
11. Marriott et al. 1999.
12. Marriott 2012.
13. Harney Peak 7,244 feet, Bear Mountain 7,164 feet, Crooks
Tower 7,137 feet, Terry Peak 7,082 feet, and Crows Nest Peak
7,045 feet. Harney Peak is in the Central Area; the others are

31. Quotation is from Dodge (1876, p. 62); photos in Progulske 1974.


32. Lovaas 1976.
33. Baker et al. (2007) concluded that a variable-severity fire model is
often more realistic and is pertinent to the restoration of forest
structure; see also Shinneman and Baker 1997.
34. Fisher et al. (1987) wrote about the effects of Native Americans
near Devils Tower.

345

346

Notes to Pages 275294


35. Brown and Cook 2006; Brown and Schoettle 2008; Brown et al.
2008.
36. Shinneman and Baker 1997; Baker et al. 2007.

9. Mason 1987 (p. 41).


10. Laramie Chamber of Commerce 1913.
11. In 2010 hay was grown on 59,000 acres, about 3 percent of the

37. Shepperd and Battaglia 2002.

Laramie Basin, and almost entirely consisted of native grasses

38. Brown 2006; Brown et al. 2008.

(no or little alfalfa). Most alfalfa in Wyoming is grown in

39. Orr 1975.

Goshen and Fremont counties. Data from U.S. Department of

40. Early journals reviewed by Shepperd and Battaglia (2002).


41. Lovaas 1976; Wind Cave National Park has used prescribed fires
since the early 1970s.

Agriculture National Agricultural Statistics Service.


12. Mears 1991, 2001.
13. The annual water supply for the city of Laramie is approxi-

42. Battaglia et al. 2008.

mately 60 percent from surface water draining from the Medi-

43. Brown (2006) found that fires were more frequent during La

cine Bow Mountains of Wyoming and Colorado, and 40 percent

Nia years, cool phases of the Pacific Decadal Oscillation, and


warm phases of the Atlantic Multidecadal Oscillation.
44. Lentile et al. 2006.

from wells into the Casper aquifer.


14. The Casper Aquifer Protection Plan is at http://www.ci.laramie
.wy.us/index.aspx?NID=226.

45. Shepperd and Battaglia (2002) review the ecology of the vari-

15. Mears 1991; katabatic winds are generated by air becoming

ous insects and fungal diseases common in the area, namely,

cooler and denser on glacier surfaces, then flowing down

mountain pine beetle, pine engraver, red turpentine beetle,

glacial and mountain slopes; they are well known in Ant-

Armillaria root rot, red rot, western gall rust, needle cast, and

arctica and New Zealand. See http://en.wikipedia.org/wiki/

diplodia tip blight; see also Negron et al. 2008; chapters 11


and 12.

Katabatic_wind.
16. The Wyoming Water Resources Data System (http://www

46. The most recent condition assessments are available online or

.wrds.uwyo.edu/) provided temperature and precipitation data

from the headquarters of the parks or Devils Tower National

for a 2.5 mile 2.5 mile area in the Laramie Basin (41.25N,

Monument.
47. Battaglia et al. (2008) report that fires have helped reduce some
exotic plants, and that 39 percent of the grasslands and 19 percent of the forests in the Black Hills had prescribed burns in the
previous 10 years.
48. Wind Cave National Parks ferret population was estimated to
be 4664 in 2013; see the parks website for current estimates.
49. Pauli et al. 2006.
50. B. Burkhart, pers. comm.
51. Pronghorn prefer forbs, which are more common in prairie dog

105.75W), using PRISM Climate Group data, Oregon State University (http://www.prism.oregonstate.edu/terms.phtml).
17. Irrigation developments have increased the size of wetlands in
the basins.
18. Black sagebrush is found just east of Laramie, where Grand
Avenue joins Interstate 80; exposed bedrock in the same area is
dominated by true mountain-mahogany; see Thatcher 1959.
19. Elliott-Fisk et al. 1983; Norris et al. 2006; S. T. Jackson, pers.
comm.
20. Peck and Lovvorn 2001; Sueltenfuss et al. 2013.

colonies; bison graze on the younger, less fibrous grass sprouts

21. The Wyoming Game and Fish Department has a report titled

that are easily accessible in the colonies because of the clipping

Laramie Plains wetland complex: Regional wetland conserva-

done by the rodents.


52. Rehfeldt et al. 2006; see also Bartlein et al. 1997.
53. Shepperd and Battaglia 2002; King et al. 2013.

tion plan (March 19, 2012).


22. Beath et al. 1941; Kolm 1975; there are two groups of plants
that accumulate selenium. One, the selenium indicators,
require selenium for their growth (two-grooved milkvetch,

Chapter 17: The Laramie Basin

tineleaved milkvetch, woody aster, and princesplume). In


contrast, the other comprises facultative accumulators, such

1. The one basin with no outlet is the Great Divide Basin, where

as Gardners saltbush, gumweed, locoweed, mentzelia, snake-

the Continental Divide splits for a distance of about 100 miles

weed, spiny hopsage, winterfat, and certain grasses. They do

(see fig. 1.2).

not require selenium but often develop high concentrations of

2. Prasciunas et al. 2008.

the element. Some plants do not accumulate selenium, even

3. During the Pleistocene, much of the Laramie Basin was a cold,

though they are frequently found on seleniferous soils. Both

semi-arid, tundra-like steppe, as indicated by the presence of

selenium indicators and facultative accumulators can concen-

fossil ice wedges; see chapter 2.

trate large amounts of the element in leaves and stems without

4. Shuman 2012.

adverse effects on the plant itself. For example, two-grooved

5. Kornfeld et al. 2010.

milkvetch is known to accumulate 4,000 ppm or more from

6. Homsher 1965.

soils containing only 1.1 ppm in the soil. Grasses and cereal

7. Burdick 1987 (pp. 23); also, Homsher (1965) and Mason (1987)

grains with only 1030 ppm can be lethal or crippling to cattle

provide overviews of human history in the Laramie Basin.

and horses. Both the indicators and the accumulators increase

8. Fremont 1845; Stansbury 1852; Smith 1913 (p. 272); Dale 1918

surface-soil selenium concentrations as their shoots decom-

(letter from Ashley to Atkinson Dec. 1, 1825); Parkman 1950;

pose in the fall, thereby possibly improving conditions for

see also Dorn 1986.

their own growth.

Notes to Pages 295320


23. Myers 1969.
24. M. T. Richardson (unpublished) reports that construction of the

disruption of disturbance regimes; it can be viewed at http://


wgfd.wyo.gov/web2011/wildlife-1000407.aspx.

Pioneer Canal from the Big Laramie River to Sodergren Lake

5. http://www.rivers.gov/rivers/wyoming.php.

was begun in 1879. This canal was extended to Lake Hattie in

6. Ashworth 2006.

1903. Units 2 and 3 of the Wheatland Reservoirs were created

7. Curtis and Grimes 2004 (p. 106); www.wrds.uwyo.edu.

by damming the Laramie River in the north end of the Laramie

8. Kornfeld and Osborn 2003.

Basin (see fig. 17.1). Richardson presents anecdotal evidence

9. Baker and Knight 2000; Ingelfinger and Anderson (2004) found

suggesting that irrigation caused an increase in the abundance

that the densities of Brewers sparrow and sage sparrow were

of mosquitos. For a map of the Pioneer CanalLake Hattie Irri-

reduced by 3960 percent within 300 feet of roads having low

gation District, see http://waterplan.state.wy.us/plan/platte/

traffic. See also Sawyer et al. (2007, 2009) for studies on the

atlas/upper/upper_agricultural_irrigation_districts.htm.
25. Lovvorn et al. 1999; Peck and Lovvorn 2001.
26. Laramie Sentinel, June 7, 1879; Laramie Boomerang, 1892 (interview with W. O. Owen).

effects of roads and traffic on elk and mule deer.


10. Ostlund 2012; prior to the completion of the Trappers Point
project, about 700 pronghorn and mule deer were killed along
a 27-mile stretch of U.S. Highway 191 during a 5-year period.

27. Peck et al. 2004; Shinker et al. 2010.

11. Pocewicz et al. 2011.

28. For a summary of Wyoming water law, see http://library.wrds

12. Knight and Reiners 2000.

.uwyo.edu/wrp/90-17/90-17.html. Water rights are granted


according to the first in time, first in right doctrine.
29. For weeds and pests, see http://www.wyoweed.org/statelist.html.

13. Kuvlesky et al. 2007; Pruett et al. 2009; Knick and Connelly
2011; Dinkens et al. 2012; LeBeau et al. 2014; Hamerlinck et al
(2013) describe recent trends in land use.

For fish- and game-related nuisance species, see http://wgfd

14. Eminent ecologist Eugene Odum suggested that a region could

.wyo.gov/web2011/Departments/Hunting/pdfs/Regulations_

be divided into four fundamental kinds of landscapes: urban,

Ch62.pdf.

industrial, agricultural, and protective. Properly managed

30. Ranchers prefer more grasses than sedges and rushes, which

rangelands can be considered as protective (see chapter 6).

tend to increase with excess irrigation; slender wheatgrass,

15. Hansen et al. 2002 (p. 160); see also Power 2005.

thickspike wheatgrass, and western wheatgrass are sometimes

16. Bormann 1976 (p. 759).

common along the edges of irrigated meadows.

17. Daily and Ellison 2002; Clark 2007; Esty 2007; Banerjee et al.

31. M. Curry, pers. comm.


32. Data from the Wyoming State Climate Office.
33. See the following website for other comparisons: http://www
.wrds.uw yo.edu/sco/data/normals/1971-2000/delta-Precip
-Annual.html.
34. Reiners 2003.
35. Norris et al. 2006.

2013; Hansen et al. 2013; the 2008 Farm Bill called for measuring the environmental benefits of conservation.
18. Hansen et al. 2002; Copeland et al. 2007, 2009; Kiesecker et al.
2007, 2009a,b; Rissman et al. 2007.
19. Pocewicz et al. (2013) identified specific areas important for
migratory birds.
20. Naugle 2011; Jakle 2012a,b.
21. A safe harbor agreement assures landowners that the U.S.

Chapter 18: Using Western Landscapes


1. Larson 1977.
2. Unruh 1993.
3. Madson 2007 (p. 4); see also Madson 2006b. Worster (1994)
wrote about the New West as an unsettled country. Baron
et al. (2013) summarize the rapid, unexpected changes that
occurred from 2002 to 2012 in the Rocky Mountain region;
Hamerlinck et al. (2013) describe land use trends in Wyoming.
4. The Wyoming Game and Fish Department developed a 900page plan for the conservation of wildlife and the habitat
they depend on. The plan describes the status of individual
species and identifies five primary threats: rural subdivision,
energy development, invasive species, climate change, and the

Fish and Wildlife Service will not require different management activities without the landowners consent, and that, at
the end of the agreement period, participants may return the
property to the conditions that existed at the beginning of the
agreement.
22. Middleton et al. 2013b,c,d.
23. Williams and Jackson 2007; Williams et al. 2007; Breshears et
al. (2005) describe global-change-type drought that has greatly
reduced the abundance of pinyon pine in the southern Rocky
Mountains.
Epilogue
1. Leopold 1938 (p. 254).

347

This page intentionally left blank

References

Aanderud, Z. T. and J. H. Richards. 2009. Hydraulic redistribution


may stimulate decomposition. Biogeochem. 95:32333.
Abbott, E. C., and H. H. Smith. 1955. We Pointed Them North.
Univ. Oklahoma Press, Norman.
Ahlbrandt, T. S. 1973. Sand dunes, geomorphology and geology,
Killpecker Creek area, northern Sweetwater County, Wyoming.
M.S. thesis, Univ. Wyoming, Laramie.
Akashi, Y. 1988. Riparian vegetation dynamics along the Bighorn
River, Wyoming. M.S. thesis, Univ. Wyoming, Laramie.
Albertson, F. W., and J. E. Weaver. 1944. Nature and degree of recovery of grassland from the great drought of 1933 to 1940.
Ecol. Monogr. 14:393479.
Alexander, R. R. 1986. Silvicultural systems and cutting methods
for old-growth lodgepole pine forests in the central Rocky
Mountains. U.S. Forest Ser. Gen. Tech. Rep. RM-127.
. 1987a. Silvicultural systems, cutting methods, and cultural
practices for Black Hills ponderosa pine. U.S. Forest Ser. Gen.
Tech. Rep. RM-139.
. 1987b. Ecology, silviculture, and management of the Engelmann sprucesubalpine fir type in the Central and Southern
Rocky Mountains. U.S. Forest Ser. Agric. Handbook 659.
Allan, J. D., and A. S. Flecker. 1993. Biodiversity conservation in
running waters. BioScience 43:3243.
Allen, E. B., and D. H. Knight. 1984. The effects of introduced annuals on secondary succession in sagebrush-grassland, Wyoming. Southwest. Nat. 29:40721.
Allen, M. F. 1991. The Ecology of Mycorrhizae. Cambridge Univ.
Press, Cambridge.

Allred, B. W. 1941. Grasshoppers and their effect on sagebrush on


the Little Powder River in Wyoming and Montana. Ecology
22:38792.
Amman, G. D., compiler. 1989. Symposium on the management of
lodgepole pine to minimize losses to the mountain pine beetle.
U.S. Forest Ser. Gen. Tech. Rep. INT-262.
Amman, G. D., and K. C. Ryan. 1991. Insect infestation of fireinjured trees in the Greater Yellowstone Area. U.S. Forest Ser.
Res. Note INT-398.1.
Amundson, M. A. 1991. Wyoming: Time and Again. Pruett Publishing, Boulder.
Anderegg, W. R. L., J. W. Prall, J. Harold, and S. H. Schneider. 2010.
Expert credibility in climate change. Proc. Natl. Acad. Sci. USA
107:121079.
Anderegg, W. R. L., L. Plavcova, L. D. L. Anderegg, et al. 2013.
Droughts legacy: Multiyear hydraulic deterioration underlies
widespread aspen forest die-off and portends increased future
risk. Global Change Biol. 19:118896.
Anderson, E. M. 2007. Changes in bird communities and associated
habitats associated with fed elk. Wilson J. Ornith. 119:400409.
Anderson, J. E., and K. E. Holte. 1981. Vegetation development over
twenty-five years without grazing on sagebrush-dominated
rangeland in southeastern Idaho. J. Range Manage. 34:2529.
Anderson, J. E., and R. S. Inouye 2001. Landscape-scale changes
in plant species abundance and diversity of a sagebrush steppe
over 45 years. Ecol. Monogr. 71:53156.
Anderson, M. D., and W. L. Baker. 2005. Reconstructing landscapescale tree invasion using survey notes in the Medicine Bow
Mountains, Wyoming, USA. Landscape Ecol. 21:24358.

Allen, M. F., and J. A. MacMahon. 1985. Impact of disturbance on

Aoki, C. F., W. H. Romme, and M. E. Rocca. 2011. Lodgepole pine

cold desert fungi: Comparable microscale dispersion patterns.

seed germination following tree death from mountain pine

Pedobiologia 28:21524.
Allen, M. F., E. B. Allen, and N. E. West. 1987. Influence of parasitic

beetle attack in Colorado, USA. Am. Midl. Nat. 165:44651.


Aplet, G. H., R. D. Laven, and F. W. Smith. 1988. Patterns of com-

and mutualistic fungi on Artemisia tridentata during high pre-

munity dynamics in Colorado Engelmann sprucesubalpine fir

cipitation years. Bull. Torrey Bot. Club 114:27279.

forests. Ecology 69:31219.

349

350 References
Archer, S., M. G. Garrett, and J. K. Detling. 1987. Rates of vegeta-

Baker, W. L., and D. J. Shinneman. 2004. Fire and restoration of

tion change associated with prairie dog (Cynonomys ludovidana)

pinon-juniper woodlands in the western United States: A re-

grazing in North American mixed-grass prairie. Vegetatio


72:15966.
Arkley, R. J. and H. C. Brown. 1954. The origin of mima mound
(hogwallow) microrelief in the far western states. Soil Sci. Soc.
Am. Proc. 18:19599.
Armstrong, D. M., J. C. Halfpenny, and C. H. Southwick. 2001.
Vertebrates. In W. D. Bowman and T. R. Seastedt, eds., Structure
and Function of an Alpine Ecosystem, pp. 12856. Oxford Univ.
Press, New York.
Arno, S. F., and G. E. Gruell. 1986. Douglas fir encroachment into
mountain grasslands in southwestern Montana. J. Range Manage. 39:27276.
Arno, S. F., and R. P. Hammerly. 1984. Timberline: Mountain and
Arctic Forest Frontiers. Mountaineers, Seattle.
Arno, S. F., and A. E. Wilson. 1986. Dating past fires in curlleaf
mountain-mahogany communities. J. Range Manage. 39:
24143.
Aschan, G., C. Wittmann, and H. Pfanz. 2001. Age-dependent bark
photosynthesis of aspen twigs. Trees 15:43137.
Ashworth, W. 2006. Ogallala Blue. Countryman Press, Woodstock,
Vermont.
Axelrod, D. I. 1985. Rise of the grassland biome, central North
America. Bot. Rev. 51:163201.
Baker, B. W., H. C. Ducharme, D. C. S. Mitchell et al. 2005. Inter
action of beaver and elk herbivory reduces standing crop of willow. Ecol. Appl. 15:11018.
Baker, R. G. 1986. Sangamonian (questionable) and Wisconsinan
paleoenvironments in Yellowstone National Park. Geol. Soc.
Am. Bull. 97:71736.
Baker, W. L. 1989a. Classification of the riparian vegetation of the
montane and subalpine zones in western Colorado. Great Basin
Naturalist 49:21428.
. 1989b. Macro- and micro-scale influences on riparian
vegetation in western Colorado. Ann. Assoc. Am. Geogr. 79:
6578.
. 1990. Climatic and hydrologic effects on the regeneration

view. Forest Ecol. Manage. 189:121.


Baker, W. L., and T. T. Veblen. 1990. Spruce beetles and fires in
the nineteenth-century subalpine forests of western Colorado,
USA. Arct. Alp. Res. 22:6580.
Baker, W. L., and P. J. Weisberg. 1995. Landscape analysis of the
forest-tundra ecotone in Rocky Mountain National Park, Colorado. Prof. Geographer 47:36175.
. 1997. Using GIS to model tree population parameters in
the Rocky Mountain National Park forest-tundra ecotone.
J.Biogeogr. 24:51326.
Baker, W. L., T. T. Veblen, and R. L. Sherriff. 2007. Fire, fuels and
restoration of ponderosa pine-Douglas-fir forests in the Rocky
Mountains, USA. J. Biogeogr. 34:25169.
Bamberg, S. A. 1971. Plant ecology of alpine tundra areas in Montana
and adjacent Wyoming. M.A. thesis, Univ. Colorado, Boulder.
Banerjee, S., S. Secchi, J. Fargione, et al. 2013. How to sell ecosystem
services:a guide for designing new markets. Frontiers Ecol. Env.
11:297304.
Baril, L. M., A. J. Hansen, R. Renkin, et al. 2011. Songbird response
to increased willow (Salix spp.) growth in Yellowstones northern range. Ecol. Appl. 21:228396.
Barker, J. R., and C. M. McKell. 1986. Differences in big sagebrush
(Artemisia tridentata) plant stature along soil-water gradients:
Genetic components. J. Range Manage. 39:14751.
Barnes, P. W., and A. T. Harrison. 1982. Species distribution and
community organization in a Nebraska Sandhills mixed prairie
as influenced by plant/soil water relationships. Oecologia 52:
192201.
Barnosky, C. W. 1984. Late Miocene vegetational and climatic variations inferred from a pollen record in northwest Wyoming.
Science 223:4951.
Baron, J. S., and D. H. Campbell. 1997. Nitrogen fluxes in a high
elevation Colorado Rocky Mountains basin. Hydrol. Processes
11:78399.
Baron, J. S., D. S. Ojima, E. A. Holland, and W. J. Parton. 1994.
Analysis of nitrogen saturation potential in Rocky Moun-

of Populus angustifolia James along the Animas River, Colorado.

tains tundra and forestImplication for aquatic systems. Bio

J. Biogeogr. 17:5973.

geochemistry 27:6182.

. 2006. Fire and restoration of sagebrush ecosystems. Wildlife Soc. Bull. 34:17785.
. 2009. Fire Ecology in Rocky Mountain Landscapes. Island
Press, Washington, D.C.
. 2011. PreEuro-American and recent fire in sagebrush ecosystems. In S. T. Knick and J. W. Connelly, eds., pp. 185201,
Greater Sage-grouse: Ecology and Conservation of a Landscape
Species and Its Habitats. Univ. California Press, Berkeley.
Baker, W. L., and S. C. Kennedy. 1985. Presettlement vegetation of
part of northwestern Moffat County, Colorado, described from
remnants. Great Basin Nat. 45:74777.
Baker, W. L., and R. L. Knight. 2000. Roads and forest fragmenta-

Baron, J. S., T. Seastedt, D. Fagre, et al. 2013. Meeting Reports:


Rocky Mountain futures: Preserving, utilizing, and sustaining
Rocky Mountain ecosystems. Bull. Ecol. Soc. Am. 94:19599.
Bartlein, P. J., C. Whitlock, and S. I. Shafer. 1997. Future climate in
the Yellowstone National Park region and its potential impact
on vegetation. Conserv. Biol. 11:78292.
Bartos, D. L., and G. D. Amman. 1989. Microclimate: An alternative
to tree vigor as a basis for mountain pine beetle infestations.
U.S. Forest Ser. Res. Paper INT-400.
Bartos, D. L., W. F. Mueggler, and R. B. Campbell, Jr. 1991. Regeneration of aspen by suckering on burned sites in western Wyoming. U.S. Forest Ser. Res. Paper INT-448.

tion in the Southern Rockies. In R. L. Knight, F. W. Smith, S. W.

Bartz, J. E. M. 1997. Playa plant community patterns and composi-

Buskirk, et al., eds., Forest Fragmentation in the Southern Rocky

tions in the Powder River Basin of northeast Wyoming. M.S.

Mountains, pp. 97122. Univ. Press Colorado, Boulder.

thesis, Univ. Wyoming, Laramie.

References
Bates, J. D., T. Svejcar, R. F. Miller, and R. A. Angell. 2006. The ef-

Benkman, C. W., A. M. Siepielski, and J. W. Smith. 2012. Con-

fects of precipitation timing on sagebrush steppe vegetation.

sequences of trait evolution in a multi-species system. In

J.Arid Environ. 64:67097.

T.Ohgushi, O. J. Schmitz, and R. D. Holt, eds., Trait-mediated

Battaglia, M. A., F. W. Smith, and W. D. Shepperd. 2008. Can prescribed fire be used to maintain fuel treatment effectiveness
over time in Black Hills ponderosa pine forests? Forest Ecol.
Manage. 256:202938.
Beath, O. A., C. S. Gilbert, and H. F. Eppson. 1941. The use of indicator plants in locating seleniferous areas in western United
States. Am. J. Bot. 28:887900.

Indirect Interactions: Ecological and Evolutionary Perspectives,


pp. 27891. Cambridge Univ. Press, Cambridge.
Bentz, B. J., J. Regniere, C. J. Fettig, et al. 2010. Climate change and
bark beetles of the western United States and Canada: Direct
and indirect effects. BioScience 60:60213.
Berg, A. W. 1990. Formation of mima mounds: A seismic hypothesis. Geology 18:28184

Beauvais, G. P. 2008. Vertebrate wildlife of the Red Desert. In A.

Berger, J., P. B. Stacey, L. Bellis, and M. P. Johnson. 2001. A mam

Proulx, ed., Red Desert: History of a Place, pp. 14370. Univ.

malian predator-prey imbalance: Grizzly bear and wolf extinc-

Texas Press, Austin.

tion affect avian neotropical migrants. Ecol. Appl. 11:94760.

Bechtold, H. A., and R. S. Inouye. 2007. Distribution of carbon and

Berger, K. M., E. M. Gese, and J. Berger. 2008. Indirect effects and

nitrogen in sagebrush steppe after six years of nitrogen addition

traditional trophic cascades: A test involving wolves, coyotes,

and shrub removal. J. Arid Environ. 71:12232.

and pronghorn. Ecology 89:81828.

Beck, J. L., J. W. Connelly, and K. P. Reese. 2009. Recovery of greater

Bergmann, D., M. Zehfus, L. Zierer, et al. 2009. Grass rhizosheaths:

sage-grouse habitat features in Wyoming big sagebrush follow-

Associated bacterial communities and potential for nitrogen

ing prescribed fire. Restoration Ecol. 17:393403.

fixation. West. N. Am. Nat. 69:10514.

Beck, J. L., J. W. Connelly, and C. L. Wamboldt. 2012. Conse-

Bernal, B., and W. J. Mitsch. 2012. Comparing carbon sequestration

quences of treating Wyoming big sagebrush to enhance wildlife

in temperate freshwater wetland communities. Global Change

habitats. Rangeland Ecol. Manage. 65:44455.


Beetle, A. A. 1974a. Range survey in Teton County, Wyoming. Part
4, Quaking aspen. Wyoming Agric. Exp. Sta. Sci. Monogr. 27,
Laramie.
. 1974b. The zootic disclimax concept. J. Range Manage. 27:
3032.
Beetle, A. A., and K. L. Johnson. 1982. Sagebrush in Wyoming.
Wyoming Agric. Exp. Sta. Bull. 779, Laramie.
Beever, E. 2003. Management implications of the ecology of freeroaming horses in the semi-arid ecosystems of the western
United States. Wildlife Soc. Bull. 31:88795.
Behan, M. J. 1957. The vegetation and ecology of Dry Park in the

Biol. 18:163647.
Beschta, R. L. 2003. Cottonwoods, elk, and wolves in the Lamar
Valley of Yellowstone National Park. Ecol. Appl. 13:12951309.
. 2005. Reduced cottonwood recruitment following extirpation of wolves in Yellowstones northern range. Ecology 86:
391403.
Beschta, R. L., and W. J. Ripple. 2010. Recovering riparian plant
communities with wolves in Northern Yellowstone, USA. Restoration Ecology 18:38089.
Betancourt, J. L., W. S. Schuster, J. B. Mitton, and R. S. Anderson.
1991. Fossil and genetic history of a pinyon pine (Pinus edulus)
isolate. Ecology 72:168597.

Medicine Bow Mountains. M.S. thesis, Univ. Wyoming, Laramie.

Beyer, H. L., E. H. Merrill, N. Varley, and M. S. Boyce. 2007. Willow

Beiswenger, J. M. 1991. Late Quaternary vegetational history of

on Yellowstones northern range: Evidence for a trophic cas-

Grays Lake, Idaho. Ecol. Monogr. 6:16582.


Bekker, M. F., and G. P. Malanson. 2008. Linear forest patterns in
subalpine environments. Prog. Phys. Geog. 32:63553.
Bekker, M. F., J. T. Clark, and M. W. Jackson. 2009. Landscape metrics
indicate differences in patterns and dominant controls of ribbon

cade? Ecol. Appl. 17:156371.


Biederman, J. A., P. D. Brooks, A. A. Harpold, et al. 2012. Multiscale observations of snow accumulation and peak snowpack
following widespread, insect-induced lodgepole pine mortality.
Ecohydrology 7:15062.

forests in the Rocky Mountains, USA. Appl. Veg. Sci. 12:273349.

Bienen, L., and G. Tabor. 2006. Applying an ecosystem approach

Bell, K. L., and L. C. Bliss. 1979. Autecology of Kobresia bellardii:

to brucellosis control: Can an old conflict between wildlife

Why winter snow accumulation limits local distribution. Ecol.

and agriculture be successfully managed? Front. Ecol. Environ.

Monogr. 49:377402.

4:31927.

Belnap, J., and O. Lange 2001. Biological soil crusts: Structure, func-

Billings, W. D. 1954. Temperature Inversions in the Pinyon-juniper

tion, and management. Ecol. Studies, vol. 150. Springer, New

Zone of a Nevada Mountain Range. Butler Univ. Bot. Studies 12.

York.
Benedict, J. B. 1984. Rates of tree-island migration, Colorado Rocky
Mountains, USA. Ecology 65:82023.

Butler Univ., Indianapolis, Indiana.


. 1969. Vegetational pattern near alpine timberline as affected by fire-snowdrift interactions. Vegetatio 19:192207.

Benkman, C. W., and A. M. Siepielski. 2004. A keystone selective

. 1973. Arctic and alpine vegetations: Similarities, dif-

agent? Pine squirrels and the frequency of serotiny in lodgepole

ferences, and susceptibility to disturbance. BioScience 23:

pine. Ecology 85:208287.


Benkman, C. W., R. P. Balda, and C. C. Smith. 1984. Adaptations

697704.
. 2000. Alpine vegetation. In M. G. Barbour and W. D. Bill-

for seed dispersal and the compromises due to seed predation in

ings, eds., North American Terrestrial Vegetation, 2nd edition,

limber pine. Ecology 65:63242.

pp. 53772. Cambridge Univ. Press, Cambridge.

351

352 References
Billings, W. D., and H. A. Mooney. 1959. An apparent frost hummock-sorted polygon cycle in the alpine tundra of Wyoming.
Ecology 40:1620.
Bilyeu, D. M., D. J. Cooper, and N. T. Hobbs. 2008. Water tables
constrain height recovery of willow on Yellowstones northern
range. Ecol. Appl. 18:8092.
Binkley, D., F. W. Smith, and Y. Son. 1995. Nutrient supply and
declines in leaf area and production in lodgepole pine. Can. J.
Forest Res. 25:62128.

Boggs, K., and T. Weaver. 1994. Changes in vegetation and nutrient


pools during riparian succession. Wetlands 14:98109.
Boldt, C. E., and J. L. VanDeusen. 1974. Silviculture of ponderosa
pine in the Black Hills: The status of our knowledge. U.S. Forest
Ser. Res. Paper RM-124.
Bonnicksen, T. 1989. Fire gods and federal policy. Am. Forests
95:1416 and 6668.
Boon, S. 2012. Snow accumulation following forest disturbance.
Ecohydrology 5:27985.

Biondi, F., S. D. J. Strachan, S. Mensing, et al. 2007. Radiocarbon

Bormann, F. H. 1976. An inseparable linkage: Conservation of

analysis confirms the annual nature of sagebrush growth rings.

natural ecosystems and the conservation of fossil energy. Bio

Radiocarbon 49:123140.
Black, S. H., D. Kulakowski, B. R. Noon, and D. A. DellaSala. 2013.
Do bark beetle outbreaks increase wildfire risks in the central
U.S. Rocky Mountains? Implications from recent research. Natural Areas J. 33:5965.
Blackstone, D. L., Jr. 1988. Travelers Guide to the Geology of Wyoming, 2nd edition. Bull. 67. Geological Survey of Wyoming,
Laramie.
Blackwelder, E. 1912. The Gros Ventre slide: An active earth-flow.
Geol. Soc. Am. Bull. 23:48792.
Blair, N. 1987. The history of wildlife management in Wyoming.
Wyoming Game and Fish Dept., Cheyenne.
Blaisdell, J. P., and R. C. Holmgren. 1984. Managing intermountain
rangelands: Salt desert shrub ranges. U.S. Forest Ser. Gen. Tech.
Rep. INT-163.
Bleby, T. M., A. J. Mcelrone, and R. B. Jackson. 2010. Water uptake
and hydraulic redistribution across large woody root systems to
20 m depth. Plant Cell Environ. 33:213248.

Science. 26:75460.
. 2000. On respect for nature. Northern Rockies Conservation Cooperative News 13. Jackson, Wyoming.
Boutton, T. W., A. T. Harrison, and B. N. Smith. 1980. Distribution of biomass of species differing in photosynthetic pathway
along an altitudinal gradient in southeastern Wyoming. Oecologia 45:28798.
Bowman, C. C. 1997. Characterization of playas in northeastern
Wyoming using vegetation, soils, and hydrology. M.S. thesis,
Univ. Wyoming, Laramie.
Bowman, R. A., D. M. Mueller, and W. J. McGinnies. 1985. Soil and
vegetation relationships in a central plains saltgrass meadow.
J.Range Manage. 38:32528.
Bowman, W. D., and M. C. Fisk. 2001. Primary production. In W.
D. Bowman and T. R. Seastedt, eds., Structure and Function
of an Alpine Ecosystem, pp. 17797. Oxford Univ. Press, New
York.
Bowman, W. D., J. R. Gartner, K. Holland, and M. Wiedermann.

Bleed, A. S., and C. Flowerday. 1990. An Atlas of the Sand Hills.

2006. Nitrogen critical loads for alpine vegetation and ter-

Instit ute of Agriculture and Natural Resources, Univ. of Nebra

restrial ecosystem response: Are we there yet? Ecol. Appl.

ska, Lincoln.
Blickley, J. L., D. Blackwood, and G. L. Patricelli. 2012. Experimental
evidence for the effects of chronic anthropogenic noise on abundance of greater sage-grouse at leks. Conserv. Biol. 26:46171.
Blossey, B., and R. Notzold. 1995. Evolution of increased competitive ability in invasive nonindigenous plantsA hypothesis.
J.Ecol. 83:88789.
Blumenthal, D. M., N. R. Jordan, and M. P. Russelle. 2003. Soil carbon addition controls weeds and facilitates prairie restoration.
Ecol. Appl. 13:60515.

16:118393.
Boyce, M. S. 1989. The Jackson Elk Herd: Intensive Wildlife Management in North America. Cambridge Univ. Press, New York.
Boyd, D. S., and T. J. Svejcar. 2011. The influence of plant removal
on succession in Wyoming big sagebrush. J. Arid Environ.
75:73441.
Bradley, B. A. 2009. Regional analysis of the impacts of climate
change on Cheatgrass invasion shows potential risk and opportunity. Global Change Biol. 15:196208.
. 2010. Assessing ecosystem threats from global and regional

Blumenthal, D. M., R. A. Chimner, J. M. Welker, and J. A. Morgan.

change: Hierarchical modeling of risk to sagebrush ecosystems

2008. Increased snow facilitates plant invasion in mixed-grass

from climate change, land use and invasive species in Nevada,

prairie. New Phytol. 179:44048.

USA. Ecography 33:198208.

Blumenthal, D. M., A. P. Norton, and T. R. Seastedt. 2010. Restoring

Bradley, B. A., and D. C. Marvin. 2011. Using expert knowledge to

competitors and natural enemies for long-term control of plant

satisfy data needs: Mapping invasive plant distributions in the

invaders. Rangelands 32:1620.

western United States. West. N. Am. Naturalist 71:30215.

Bock, J. H., and C. E. Bock. 1984. Effect of fires on woody vegetation

Bragg, D. C. 2000. Simulating catastrophic and individualistic large

in the pine-grassland ecotone of the southern Black Hills, USA.

woody debris recruitment for a small riparian system. Ecology

Am. Midi. Nat. 112:3542.


Bockino, N. K., and D. B. Tinker. 2012. Interactions of white pine
blister rust and mountain pine beetle in whitebark pine eco
systems in the Southern Greater Yellowstone Area. Natural
Areas J. 32:3140.

81:138394.
Bragg, T. B. 1982. Seasonal variations in fuel and fuel consumption
by fires in a bluestem prairie. Ecology 63:711.
Branson, F. A. 1985. Vegetation changes on western rangelands.
Range Monogr. 2. Soc. Range Manage., Denver.

References
Branson, F. A., and R. F. Miller. 1981. Effects of increased precipita-

Brown, J. K., and N. V. DeByle. 1989. Effects of prescribed fire on

tion and grazing management on northeastern Montana range-

biomass and plant succession in western aspen. U.S. Forest Ser.

lands. J. Range Manage. 34:310.

Res. Paper INT-412.

Branson, F. A., R. F. Miller, and I. S. McQueen. 1967. Geographic

Brown, K., A. J. Hansen, R. E. Keane, and L. J. Graumlich. 2006.

distribution and factors affecting the distribution of salt desert

Complex interactions shaping aspen dynamics in the Greater

shrubs in the United States. J. Range Manage. 20:28796.

Yellowstone Ecosystem. Landscape Ecol. 21:93351.

. 1970. Plant communities and associated soil and water fac-

Brown, M., T. A. Black, Z. Nesic, et al. 2010. Impact of mountain

tors on shale-derived soils in northeastern Montana. Ecology

pine beetle on the net ecosystem production of lodgepole pine

51:391407.

stands in British Columbia. Agri. Forest Meteorol. 150:25464.

. 1976. Moisture relationships in twelve northern desert

Brown, P. M. 2006. Climate effects on fire regimes and tree re-

shrub communities near Grand Junction, Colorado. Ecology

cruitment in Black Hills ponderosa pine forests. Ecology 87:

57:110424.

25002510.

Breshears, D. D., N. S. Cobb, P. M. Rich, et al. 2005. Regional vegeta-

Brown, P. M., and B. Cook. 2006. Early settlement forest structure

tion die-off in response to global-change-type drought. Proc.

in Black Hills ponderosa pine forests. Forest Ecol. Manage. 223:

Natl. Acad. Sci. USA 102:2529.

28490.

Briggs, G. M., and J. A. MacMahon. 1983. Alpine and subalpine wet-

Brown, P. M., and A. W. Schoettle. 2008. Fire and stand history in

land plant communities of the Uinta Mountains, Utah. Great

two limber pine (Pinus flexilis) and Rocky Mountain bristlecone

Basin Nat. 43:52330.


Briske, D. D. 1991. Developmental morphology and physiology

pine (Pinus aristata) stands in Colorado. Int. J. Wildland Fire


17:33947.

of grasses. In R. K. Heitschmidt and J. W. Stuth, eds., Grazing

Brown, P. M., and R. Wu. 2005. Climate and disturbance forcing

Management: An Ecological Perspective, pp. 85108. Timber

of episodic tree recruitment in a southwestern ponderosa pine

Press, Portland, Oregon.

landscape. Ecology 86:303038.

Briske, D. D., S. D. Fuhlendorf, and F. E. Smeins. 2005. State-and-

Brown, P. M., M. R. Kaufmann, and W. D. Shepperd. 1999. Long-

transition models, thresholds, and rangeland health: A synthe-

term, landscape patterns of past fire events in montane ponder-

ses of ecological concepts and perspectives. Rangeland Ecol.


Manage. 58:110.
. 2006. A unified framework for assessment and application
of ecological thresholds. Rangeland Ecol. Manage. 59:22536.
Briske, D. D., N. F. Sayre, L. Huntsinger, et al. 2011. Origin, persistence, and resolution of the rotational grazing debate: Inte-

osa pine forest of central Colorado. Landscape Ecol. 14:51332.


Brown, P. M., C. L. Wienk, and A. J. Symstad. 2008. Fire and forest
history at Mount Rushmore. Ecol. Appl. 18:198499.
Brown, R. W. 1971. Distribution of plant communities in southeastern Montana badlands. Am. Midl. Nat. 85:45877.
Brunelle, A., G. E. Rehfeldt, B. Bentz, and A. S. Munson. 2008.

grating human dimensions into rangeland research. Rangeland

Holocene records of Dendroctonus bark beetles in high eleva-

Ecol. Manage. 64:32534.

tion pine forests of Idaho and Montana, USA. Forest Ecol. Man-

Brooks, P. D., and M. W. Williams. 1999. Snowpack controls on nitrogen cycling and export in seasonally snow-covered catchments. Hydrol. Process. 13:217790.
Brookshire, E. N. J., J. B. Kauffman, D. Lytjen, and N. Otting. 2002.
Cumulative effects of wild ungulate and livestock herbivory on
riparian willows. Oecologia 132:55966.
Brotherson, J. D., D. L. Anderson, and L. A. Szyska. 1984. Habitat
relations of Cercocarpus montanus (true mountain mahogany) in
central Utah. J. Range Manage. 37:32124.
Brough, J. S. 1996. Essential hydrological functions of playas in
the Powder River Basin of northeastern Wyoming. M.S. thesis,
Univ. Wyoming, Laramie.
Brown, D. J., W. A. Hubert, and S. H. Anderson. 1996. Beaver ponds
create wetland habitat for birds in mountains of southeastern
Wyoming. Wetlands 16:12733.

age. 255:83646.
Buckner, D. L. 1977. Ribbon forest development and maintenance
in the central Rocky Mountains of Colorado. Ph.D. diss., Univ.
Colorado, Boulder.
Bull, E. L., C. G. Parks, and T. R. Torgersen. 1997. Trees and logs
important to wildlife in the Interior Columbia River Basin. U.S.
Forest Ser. PNW-GTR-391.
Bunting, D. P., S. A. Kurc, and M. R. Grabau. 2011. Using existing agricultural infrastructure for restoration practices: Factors
influencing successful establishment of Populus fremontii over
Tamarix ramosissima. J. Arid Environ. 75:85160.
Burdick, L. 1987. Overview. In M. K. Mason, ed., LaramieGem
City of the Plains, pp. 23. Curtis Media, Dallas.
Burke, I. C., W. A. Reiners, D. L. Sturges, and P. A. Matson. 1987.
Herbicide treatment effects on properties of mountain big sage-

Brown, J. F. 1997. Effects of experimental burial on survival,

brush soils after fourteen years. Soil Sci. Soc. Am. J. 51:133743.

growth, and resource allocation of three species of dune plants.

Burke, I. C., W. A. Reiners, and R. K. Olson. 1989. Topographic con-

J. Ecol. 85:15158.
Brown, J. K. 1991. Should management ignitions be used in Yellow-

trol of vegetation in a mountain big sagebrush steppe. Vegetatio


84:7786.

stone National Park? In R. B. Keiter and M. S. Boyce, eds., The

Burke, I. C., W. K. Lauenroth, and D. P. Coffin. 1995. Soil organic

Greater Yellowstone Ecosystem: Redefining Americas Wilder-

matter recovery in semiarid grasslands: Implications for the

ness Heritage, pp. 13748. Yale Univ. Press, New Haven.

Conservation Reserve Program. Ecol. Appl. 5:793801.

353

354 References
Burke, I. C., W. K. Lauenroth, and D. G. Milchunas. 1997. Bio

Carman, J. G., and J. D. Brotherson. 1982. Comparisons of sites

geochemistry of managed grasslands in central North America.

infested and not infested with saltcedar (Tamarix pentandra) and

In E. A. Paul, K. Paustian, E. T. Elliott, and C. V. Cole, eds., Soil

Russian olive (Elaeagnus angustifolia). Weed Sci. 30:36064.

Organic Matter in Temperate Ecosystems, pp. 85102. CRC

Carpenter, A. T., and N. E. West. 1987. Indifference to mountain big

Press, Boca Raton, Florida.


Burke, I. C., W. K. Lauenroth, M. F. Antolin, et al. 2008. The future
of the shortgrass steppe. In W. K. Lauenroth and I. C. Burke,

sagebrush growth to supplemental water and nitrogen. J. Range


Manage. 40:44851.
Carrera, P. E., D. A. Trimble, and M. Ruben. 1991. Holocene treeline

eds., Ecology of the Shortgrass Steppe, pp. 484503. Oxford

fluctuations in the northern San Juan Mountains, Colorado,

Univ. Press, Oxford.

USA, as indicated by radiocarbon-dated conifer wood. Arct. Alp.

Burkhardt, J. W., and E. W. Tisdale. 1976. Causes of juniper invasion in southwestern Idaho. Ecology 57:47284.
Burnham, J. L. H., and D. L. Johnson. 2012. Mima Mounds: The
Case for Polygenesis and Bioturbation. Special Paper 490. Geol.
Soc. Am., Boulder.
Burns, D. A. 2003. Atmospheric nitrogen deposition in the Rocky
Mountains of Colorado and southern WyomingA review
and new analysis of past study results. Atmos. Environ.
37:92132.

Res. 23:23346.
Carrillo, Y., F. A. Dijkstra, E. Pendall, et al. 2012. Controls over soil
nitrogen pools in a semiarid grassland under elevated CO2 and
warming. Ecosystems 15:76174.
Carter, W. A. 1939. Diary of Judge William A. Carter. Ann. Wyoming 11:75110.
Cary, L. E. 1966. A study of forest margins. M.S. thesis, Univ. Wyoming, Laramie.
Case, J. C., C. S. Arneson, and L. L. Hallberg. 1998. Preliminary

. 2004. The effects of atmospheric nitrogen deposition in the

1:500,000-scale digital surficial geology map of Wyoming. Geo-

Rocky Mountains of Colorado and southern Wyoming, USAA

logical Hazards Section Digital Map 98-1. Geological Survey of

critical review. Environ. Pollution 127:25769.

Wyoming, Laramie.

Bush, K. L., C. K. Dyte, B. J. Moynahan, et al. 2011. Population

Cawker, K. B. 1980. Evidence of climatic control from population

structure and genetic diversity of greater sage-grouse (Centro-

age structure of Artemisia tridentata Nutt. in southern British

cercus urophasianus) in fragmented landscapes at the northern


edge of their range. Conserv. Genet. 12:52742.

Columbia. J. Biogeogr. 7:23748.


Chadde, S. W., and C. E. Kay. 1991. Tall willow communities on

Buskirk, M. E. 2009. Packrat. Wyoming Wildlife 73:2229.

Yellowstones northern range: A test of the natural regulation

Buskirk, S. W., L. F. Ruggiero, and C. J. Krebs. 1999. Habitat frag-

paradigm. In R. R. Keiter and M. S. Boyce. eds., The Greater

mentation and interspecific competition: Implications for lynx

Yellowstone Ecosystem: Redefining Americas Wilderness Heri-

conservation. In L. F. Ruggiero et al., eds., Ecology and conser-

tage, pp. 23162. Yale Univ. Press, New Haven.

vation of lynx in the United States, pp. 83100. U.S. Forest Ser.
RMRS GTR 30WWW.
Butler, D. R. 1985. Vegetation and geomorphic change on snow avalanche paths, Glacier National Park, Montana, USA. Great Basin
Nat. 45:31317.
Buttrick, P. L. 1914. The probable origin of the forests of the Black
Hills of South Dakota. Forest Quart. 12:22327.
Byers, D. A., and C. S. Smith. 2007. Ecosystem controls and the
archaeofaunal record: An example from the Wyoming Basin,
USA. Holocene 17:117183.
Caldwell, M. M., T. E. Dawson, and J. H. Richards. 1998. Hydraulic lift: Consequences of water efflux from the roots of plants.
Oecologia 113:15161.
Call, M. W. 1966. Beaver pond ecology and beaver-trout relationships in southeastern Wyoming. Ph.D. diss., Univ. Wyoming,
Laramie.
Camp, C. L., ed. 1960. James Clyman, Frontiersman, 17921881:

Chadde, S. W., P. L. Hansen, and R. D. Pfister. 1988. Wetland Plant


Communities of the Northern Range, Yellowstone National
Park. School of Forestry, Univ. Montana, Missoula.
Chadde, S. W., J. Shelly, R. J. Bursik, et al. 1998. Peatlands on national forests of the Northern Rocky Mountains: Ecology and
conservation. U.S. Forest Ser. RMRS-GTR-11.
Chadwick, H. W., and P. D. Dalke. 1965. Plant succession on dune
sands in Fremont County, Idaho. Ecology 46:76580.
Chambers, J. C., and B. E. Norton. 1993. Effects of grazing and
drought on population dynamics of salt desert shrub species
on the Desert Experimental Range, Utah. J. Arid Environ.
24:26175.
Chambers, J. C., B. A. Roundy, R. R. Blank, et al. 2007. What makes
Great Basin sagebrush ecosystems invasible by Bromus tectorum?
Ecol. Monogr. 77:11745.
Chapin, F. S., III, P. A. Matson, and P. Vitousek. 2012. Principles of
Terrestrial Ecosystem Ecology. Springer, London.

The Adventures of a Trapper and Covered Wagon Immigrant as

Chapman, T. B., T. T. Veblen, and T. Schoennagel. 2012. Spatiotem-

Told in His Own Reminiscences and Diaries. Champoeg Press,

poral patterns of mountain pine beetle activity in the southern

Cleveland.

Rocky Mountains. Ecology 93:217585.

Campbell, J. S. 1997. North American alpine ecosystems. In F. E.

Charley, J. L. 1977. Mineral cycling in rangeland ecosystems. In

Wielgolaski, ed., Ecosystems of the World: Polar and Alpine

R. E. Sosebee, ed., Rangeland Plant Physiology, pp. 21556.

Tundra, pp. 21162. Elsevier, New York.

Range Science Series 4. Soc. Range Manage., Denver.

Cannone, N., S. Sgorbati, and M. Guglielmoin. 2007. Unexpected

Charley, J. L., and N. West 1975. Plant-induced soil chemical pat-

impacts of climate change on alpine vegetation. Front. Ecol.

terns in some shrub-dominated semi-desert ecosystems of

Environ. 5:36064.

Utah. J. Ecol. 63:94563.

References
Chase, A. 1986. Playing God in Yellowstone: The Destruction of

Collins, B. J., C. C. Rhoades, R. M. Hubbard, and M. A. Battaglia.

Americas First National Park. Atlantic Monthly Press, New York.

2011. Tree regeneration and future stand development after

Choudhuri, G. N. 1968. Effect of soil salinity on germination

bark beetle infestation and harvesting in Colorado lodgepole

and survival of some steppe plants in Washington. Ecology


49:46571.
Christensen, N. L., J. K. Agee, P. F. Brossard, et al. 1989. Interpreting
the Yellowstone fires of 1988. BioScience 39:67883.
Christiansen, E., R. H. Waring, and A. A. Berryman. 1987. Resistance of conifers to bark beetle attack: Searching for general
relationships. Forest Ecol. Manage. 22:89106.
Cid, M. S., J. K. Detling, A. D. Whicker, and M. A. Brizuela. 1991.
Vegetational response of a mixedgrass prairie site following ex-

pine stands. Forest Ecol. Manage. 261:216875.


Collins, B. J., C. C. Rhoades, M. A. Battaglia, and R. M. Hubbard.
2012. The effects of bark beetle outbreaks on forest development, fuel loads and potential fire behavior in salvage logged
and untreated lodgepole pine forests. Forest Ecol. Manage.
284:260268.
Collins, E. I., R. W. Lichvar, and E. F. Evert. 1984. Description of
the only known fen-palsa in the contiguous United States. Arct.
Alp. Res. 16:25558.

clusion of prairie dogs and bison. J. Range Manage. 44:100105.

Cook, E. R., C. A. Woodhouse, C. M. Eakin, et al. 2004. Long-

Ciesla, W. 2011. Report on the Health of Colorados Forests. Colo-

term aridity changes in the western United States. Science

rado State Forest Service, Fort Collins.


Clark, F. E. 1977. Internal cycling of 15-nitrogen in shortgrass
prairie. Ecology 58:132233.
Clark, F. E., C. V. Cole, and R. A. Bowman. 1980. Nutrient cycling.
In A. I. Breymeyer and G. M. Van Dyne, eds., Grasslands, Systems Analysis and Man, pp. 659712. Cambridge Univ. Press,
Cambridge.
Clark, S. 2007. A Field Guide to Conservation Finance. Island Press,
Washington, D.C.

306:101518.
Cooke, H. A., and S. Zack. 2008. Influence of beaver dam density on
riparian areas and riparian birds in shrub steppe of Wyoming.
West. N. Am. Nat. 68:36573.
Cooke, P. G. 1857. Scenes and Adventures in the Army. Lindsay and
Blakiston, Philadelphia.
Coop, J. D., and T. Givnish. 2008. Constraints on tree seedling establishment in montane grasslands of the Valles Caldera, New
Mexico. Ecology 89:110111.

Clarke, M. L., and H. M. Rendell. 2003. Late Holocene dune accre-

Cooper, D. J., and R. E. Andrus. 1994. Patterns of vegetation and

tion and episodes of persistent drought in the Great Plains of

water chemistry in peatlands of the west-central Wind River

Northeastern Colorado. Quaternary Sci. Rev. 22:1013.


Clary, W. P., E. D. McArthur, D. Bedunah, and C. L. Wamboldt, com-

Range, Wyoming, U.S.A. Can. J. Bot. 72:158697.


Cooper, D. J., D. M. Merritt, D. C. Andersen, and R. A. Chimner.

pilers. 1992. Symposium on ecology and management of riparian

1999. Factors controlling the establishment of Fremont cotton

shrub communities. U.S. Forest Serv. Gen. Tech. Rep. INT-289.

wood seedlings on the upper Green River, USA. Regulated

Cleary, M. B., E. Pendall, and B. E. Ewers. 2010. Aboveground and


belowground carbon pools after fire in mountain big sagebrush
steppe. Rangeland Ecol. Manage. 63:18796.
Clow, D. W., C. Rhoades, J. Briggs, et al. 2011. Responses of soil and
water chemistry to mountain pine beetle induced tree mortality

Rivers Res. Manage. 15:41940.


Cooper, H. W. 1953. Amounts of big sagebrush in plant communities near Tensleep, Wyoming, as affected by grazing treatment.
Ecology 34:18689.
Coops, N. C., and R. H. Waring. 2011. A process-based approach to

in Grand County, Colorado, USA. Appl. Geochem. 26:517478.

estimate lodgepole pine (Pinus contorta Dougl.) distribution in

Cluff, G. J., R. A. Evans, and J. A. Young. 1983. Desert saltgrass seed

the Pacific Northwest under climate change. Climatic Change

germination and seedbed ecology. J. Range Manage. 36:41923.

105:31328.

Clyde, W. C. 2001. Mammalian biostratigraphy of the McCullough

Copeland, C. 2010. Wetlands: An overview of issues. Congressional

Peaks area in the northern Bighorn Basin. In P. D. Gingerich,

Research Service Reports. Paper 37. http://digitalcommons.unl

ed., Paleocene-Eocene Stratigraphy and Biotic Change in the


Bighorn and Clarks Fork Basins, Wyoming, pp. 109126. Papers
on Paleontology 33. Univ. Mich., Ann Arbor.
Coates, P. S., and D. J. Delehanty. 2010. Nest predation of greater
sage-grouse in relation to microhabitat factors and predators.
J.Wildlife Manage. 74:24048.
Coble, A. P., and T. E. Kolb. 2013. Native riparian tree establishment
along the regulated Dolores River, Colorado. West. N. Am. Nat.
73:4153.
Cochran, P. H., and C. M. Berntsen. 1973. Tolerance of lodgepole
and ponderosa pine seedlings to low night temperatures. Forest
Sci. 19:27280.
Cohen, J., and R. Stratton. 2003. Home destruction within the

.edu/crsdocs/37/.
Copeland, H. E., J. M. Ward, and J. M. Kiesecker. 2007. Assessing
tradeoffs in biodiversity, vulnerability and cost when prioritizing conservation sites. J. Conserv. Biol. 3:116.
Copeland, H. E., K. E. Doherty, D. E. Naugle, et al. 2009. Mapping
oil and gas development potential in the U. S. Intermountain
West and estimating impacts to species. PLoS ONE. 4:e7400.
Copeland, H. E., S. A. Tessman, E. H. Girvetz, et al. 2010. A geo
spatial assessment on the distribution, condition, and vulnerability of Wyomings wetlands. Ecol. Indicators 10:86979.
Coppock, D. L., and J. K. Detling. 1986. Alteration of bison/prairie
dog grazing interaction by prescribed burning. J. Wildlife Manage. 50:45255.

Hayman fire perimeter. In Graham, R. T., technical ed., Hay-

Costanza, R., R. dArge, R. de Groot, et al. 1997. The value of the

man Fire case study, pp. 26392. U.S. Forest Ser. Gen. Tech. Rep.

worlds ecosystem services and natural capital. Nature 387:

RMRS GTR-114.

25360.

355

356 References
Coughenour, M. B. 1991a. Spatial components of plant-herbivore
interactions in pastoral, ranching, and native ungulate eco
systems. J. Range Manage. 44:53042.
. 1991b. Biomass and nitrogen responses to grazing of upland
steppe on Yellowstones northern winter range. J. Appl. Ecol.
28:7182.
Coughenour, M. B., and F. J. Singer. 1991. The concept of over
grazing and its application to Yellowstones Northern Range. In

Custer, G. A. 1875. Expedition to the Black Hills. 43d Congress, 2d


sess. Senate Exec. Doc. 32.
Dahl, T. E. 1990. Wetland losses in the United States, 1780s to
1980s. U.S. Department of Interior, Fish and Wildlife Service,
Washington, D.C.
Daily, G. C., and K. Ellison. 2002. The New Economy of Nature:
The Quest to Make Conservation Profitable. Island Press, Washington, D.C.

R. B. Keiter and M. S. Boyce, eds., The Greater Yellowstone Eco-

Dale, H. C. 1918. The AshleySmith Explorations and the Discovery

system: Redefining Americas Wilderness Heritage, pp. 20930.

of a Central Route to the Pacific, 18221829, with the Original

Yale Univ. Press, New Haven.

Journals. Arthur H. Clark Co., Glendale, California.

Coupland, R. T., and G. M. Van Dyne. 1979. Systems synthesis. In

Daubenmire, R. 1943a. Soil drought as a factor determining lower

R. T. Coupland, ed., Grassland Ecosystems of the World: Analy-

altitudinal limits of trees in the Rocky Mountains. Bot. Gaz.

sis of Grasslands and Their Uses, pp. 97106. Cambridge Univ.


Press, New York.
Cox, G. W. 1984. Mounds of mystery. Natural History 93 (June):
3645.
. 1990. Comment and reply to Formation of mima mounds:
A seismic hypothesis. Geology 18:125960.
. 2012. Alpine and Montane Mima Mounds of the Western
United States. Special Paper 490. Geol. Soc. Am., Boulder.
Cox, G. W., and D. W. Allen. 1987. Soil translocation by pocket
gophers in a mima mound field. Oecologia 72:20710.
Cox, G. W., and J. Hunt. 1990. Form of mima mounds in relation to
occupancy by pocket gophers. J. Mammal. 71:9094.
Cox, G. W., C. G. Gakahu, and D. W. Allen. 1987. The small stone
content of mima mounds in the Columbia Plateau and Rocky
Mountain regions: Implications for mound origin. Great Basin
Nat. 47:60919.
Crawford, J. A., R. A. Olson, N. E. West, et al. 2004. Synthesis paper:
Ecology and management of sage-grouse and sage-grouse habitat. J. Range Manage. 57:219.
Creel, S., and D. Christianson. 2009. Wolf presence and increased
willow consumption by Yellowstone elk: Implications for trophic cascades. Ecology 90:245466.
Crist, T. O. 2008. Insect populations, community interactions, and

105:113.
. 1943b. Vegetational zonation in the Rocky Mountains. Bot.
Rev. 9:32593.
. 1954. Alpine timberlines in the Americas and their interpretation. Butler Univ. Bot. Studies 11:11936.
. 1975. Ecology of Artemisia tridentata ssp. tridentata in the
state of Washington. Northw. Sci. 49:2435.
Davidson, A. D., J. K. Detling, and J. H. Brown. 2012. Ecological
roles and conservation challenges of social, burrowing, herbivorous mammals in the worlds grasslands. Front. Ecol. Environ.
10:47786.
Davies, K. W., J. D. Bates, and R. F. Miller. 2007a. Environmental
and vegetation relationships of the Artemisia tridentata ssp.
wyomingensis alliance. J. Arid Environ. 70:47894.
. 2007b. The influence of Artemisia tridentata ssp. wyoming
ensis on microsite and herbaceous vegetation heterogeneity.
J.Arid Environ. 69:44157.
Davies, K. W., J. D. Bates, and J. J. James. 2009a. Microsite and herbaceous vegetation heterogeneity after burning Artemisia tridentata steppe. Oecologia 159:597606.
Davies, K. W., T. J. Svejcar, and J. D. Bates. 2009b. Interaction of historical and nonhistorical disturbances maintains native plant
communities. Ecol. Appl. 19:153645.

ecosystem processes in the shortgrass steppe. In W. K. Lauen-

Davies, K. W., C. S. Boyd, J. L. Beck, et al. 2011. Saving the sage-

roth and I. C. Burke, eds., Ecology of the Shortgrass Steppe,

brush sea: An ecosystem conservation plan for big sagebrush

pp.21547. Oxford Univ. Press, Oxford.

plant communities. Biol. Conserv. 144:257384.

Cross, P. C., E. K. Cole, A. P. Dobson, et al. 2010. Probable causes of

Davis, G. V. 1959. A vegetative study of three relic areas located

increasing brucellosis in free-ranging elk of the Greater Yellow

within Fort Laramie National Monument. M.S. thesis, Univ.

stone Ecosystem. Ecol. Appl. 20:27888.


Crowe, D. M. 1990. The First Century: A Hundred Years of Wildlife

Wyoming, Laramie.
Day, T. A., and J. K. Detling. 1990a. Grassland patch dynamics and

Conservation in Wyoming. Wyoming Game and Fish Depart-

herbivore grazing preference following urine deposition. Ecol-

ment, Cheyenne.

ogy 71:18088.

Cui, M. 1990. The ecophysiology of seedling establishment in subalpine conifers of the Central Rocky Mountains, USA. Ph.D.
diss., Univ. Wyoming, Laramie.
Currie, P. O., and D. L. Goodwin. 1966. Consumption of forage by
black-tailed jackrabbits on salt desert ranges of Utah. J. Wild.
Manage. 30:30411.
Curtis, J., and K. Grimes. 2004. Wyoming Climate Atlas. Wyoming
Water Resources Data System, Univ. Wyoming, Laramie.
Cushman, R. C., and S. R. Jones. 1988. The Shortgrass Prairie.
Pruett Publishing, Boulder.

. 1990b. Changes in grass leaf water relations following urine


deposition. Am. Midl. Nat. 123:17178.
De Steigure, J. E. 2011. Wild Horses of the West: History and Politics
of Americas Mustangs. Univ. Arizona Press, Tucson.
Dearing, D. 2001. Plant-herbivore interactions. In W. D. Bowman
and T. R. Seastedt, eds., Structure and Function of an Alpine
Ecosystem, pp. 26684. Oxford Univ. Press, New York.
Debinski, D. M., M. E. Jakubauskas, and K. Kindscher. 2000. Montane meadows as indicators of environmental change. Environ.
Monit. Assess. 64:21325.

References
Debinski, D. M., H. Wickham, K. Kindscher, et al. 2010. Montane
meadow change during drought varies with background hydrologic regime and plant functional group. Ecology 91:167281.

Dise, N. B. 2009. Peatland response to global change. Science 326:


81011. Available on-line at www.sciencemag.org.
Diskin, M., M. E. Rocca, K. N. Nelson, C. F. Aoki, and W. H. Romme.

DeByle, N. V. 1979. Potential effects of stable versus fluctuating elk

2011. Forest developmental trajectories in mountain pine beetle

populations in the aspen ecosystem. In M. S. Boyce and L. D.

disturbed forests of Rocky Mountain National Park, Colorado.

Hayden-Wing, eds., North American Elk: Ecology, Behavior and


Management, pp. 1319. Special Publication. Univ. Wyoming,
Laramie.

Can. J. Forest Res. 41:78292.


Dodge, R. I. 1876. The Black Hills. James Miller, New York. (Reprinted 1965 by Ross and Haines, Minneapolis, Minnesota.)

DeByle, N. V., C. D. Bevins, and W. C. Fischer. 1987. Wildfire oc-

Doering, W. R., and R. G. Reider. 1992. Soils of Cinnabar Park, Med-

currence in aspen in the interior western United States. West. J.

icine Bow Mountains, Wyoming, USA: Indicators of park origin

Appl. Forest. 2:7376.

and persistence. Arct. Alp. Res. 24:2739.

DeCant, J. P. 2008. Russian olive, Elaeagnus angustifolia, alters pat-

Doescher, P. S., R. F. Miller, J. Wang, and J. Rose. 1990. Effects of ni-

terns in soil nitrogen pools along the Rio Grande River, New

trogen availability on growth and photosynthesis of Artemisia

Mexico, USA. Wetlands 28:896904.

tridentata ssp. wyomingensis. Great Basin Nat. 50:919.

Densmore, L. 2011. Thirty years later. Wyoming Wildlife 75:1421.

Doherty, K. E., D. E. Naugle, H. E. Copeland, et al. 2011a. Energy

DePuit, E. J., and M. M. Caldwell. 1973. Seasonal pattern of net

development and conservation tradeoffs: Systematic planning

photosynthesis of Artemisia tridentata. Am. J. Bot. 60:42635.

for greater sage-grouse in their eastern range. In S. T. Knick and

Despain, D. G. 1973. Vegetation of the Big Horn Mountains, Wyo-

J. W. Connelly, eds., Greater Sage-grouse: Ecology and Con-

ming in relation to substrate and climate. Ecol. Monogr. 43:

servation of a Landscape Species and Its Habitats, pp. 50516.

32955.
. 1983. Nonpyrogenous climax lodgepole pine communities
in Yellowstone National Park. Ecol. 64:23134.
. 1990. Yellowstone Vegetation: Consequences of Environment and History in a Natural Setting. Roberts Rinehart, Boulder.
Detling, J. K., and M. I. Dyer. 1981. Evidence for potential plant
growth regulators in grasshoppers. Ecology 62:48588.

Studies in Avian Biology 38. Univ. California Press, Berkeley.


Doherty, K. E., J. L. Beck, and D. E. Naugle. 2011b. Comparing ecological site descriptions to habitat characteristics influencing
greater sage-grouse nest site occurrence and success. Rangeland
Ecol. Manage. 64:34451.
Donahue, D. L. 2010. Trampling the public trust. Environ. Affairs
37: 257316.

Detling, J. K., and L. G. Klikoff. 1973. Physiological response to

Donaldson, A. B. 1875. The Black Hills expedition. In South Dakota

moisture stress as a factor in halophyte distribution. Am. Midl.

State Historical Society. 1914, vol. 7, pp. 55480. South Dakota

Nat. 90:30718.

Historical Collections, Pierre.

Detling, J. K., and E. L. Painter. 1983. Defoliation responses of west-

Donato, D. C., B. J. Harvey, W. H. Romme, et al. 2013. Bark beetle

ern wheatgrass populations with diverse histories of prairie dog

effects on fuel profiles across a range of stand structures in

grazing. Oecologia 57:6571.


Detling, J. K., C. W. Ross, M. H. Walmsley, et al. 1981. Examination of North American bison saliva for potential plant growth
regulators. J. Chem. Ecol. 7:23946.
Detling, J. K., E. L. Painter, and D. L. Coppock. 1986. Ecotypic differentiation resulting from grazing pressure: Evidence for a likely
phenomenon. In P. J. Joss, P. W. Lynch, and O. B. Williams, eds.,
Rangelands: A Resource under Siege, pp. 43133. Second Int.
Rangelands Congress. Australian Acad. Sci., Canberra.
DeWine, J. M., and D. J. Cooper. 2008. Canopy shade and the successional replacement of tamarisk by native box elder. J. Appl.
Ecol. 45:50514.
. 2010. Habitat overlap and facilitation in tamarisk and box
elder stands: Implications for tamarisk control using native
plants. Restoration Ecol. 18:34958.
Dijkstra, F. A., D. Blumenthal, J. A. Morgan, et al. 2010. Contrasting
effects of elevated CO2 and warming on nitrogen cycling in a
semiarid grassland. New Phytol. 187:42637.
Dillon, G. K., D. H. Knight, and C. B. Meyer. 2005. Historic range of
variability for upland vegetation in the Medicine Bow National
Forest, Wyoming. U.S. Forest Ser. RMRS-GTR-139.

Douglas-fir forests of Greater Yellowstone. Ecol. Appl. 23:320.


Dorf, E. 1964. The petrified forests of Yellowstone Park. Sci. Am.
210:10714.
Dorn, R. D. 1986. The Wyoming Landscape, 18051878. Mountain
West Publishing, Cheyenne, Wyoming.
Dott, R. H., and R. L. Batten. 1981. Evolution of the Earth, 3rd edition. McGraw-Hill, New York.
Drenovsky, R. E., and J. H. Richards. 2006. Low leaf N and P resorption contributes to nutrient limitation in two desert shrubs.
Plant Ecol. 183:30514.
Drenovsky, R. E., J. J. James, and J. H. Richards. 2010. Variation
in nutrient resorption by desert shrubs. J. Arid Environ. 74:
156468.
Driese, K. L., W. A. Reiners, E. H. Merrill, and K. G. Gerow. 1997. A
digital land cover map of Wyoming, USA: A tool for vegetation
analysis. J. Veg. Sci. 8:13346.
Duncan, E. 1975. The ecology of curlleaf mountain-mahogany in
southwestern Montana with special reference to use by mule
deer. M.S. thesis, Montana State Univ., Bozeman.
Dunnewald, T. J. 1930. Grass and timber soils distribution in the
Big Horn Mountains. J. Am. Soc. Agron. 22:57786.

Dinkins, J. B., M. R. Conover, C. P. Kirol, and J. L. Beck. 2012.

Dunwiddie, P. W. 1977. Recent tree invasion of subalpine mead-

Greater sage-grouse (Centrocercus urophasianus) select nest sites

ows in the Wind River Mountains, Wyoming. Arct. Alp. Res.

and brood sites away from avian predators. Auk 129:600610.

9:39399.

357

358 References
Dzialak, M. R., D. V. Olson, S. M. Harju, et al. 2011. Identifying

Epstein, H. E., R. A. Gill, J. M. Paruelo, et al. 2002. The relative

and prioritizing greater sage-grouse nesting and brood-rearing

abundance of three plant functional types in temperate grass-

habitat for conservation in human-modified landscapes. PLoS

lands and shrublands of North and South America: Effects of

ONE. 6:e26273.
Eberle, J. J., and D. R. Greenwood. 2012. Life at the top of the green-

projected climate change. J. Biogeogr. 29:87588.


Esty, A. 2007. Banking on mitigation. Am. Scientist 95:12223.

house Eocene worldA review of the Eocene flora and verte-

Evans, R. K., R. Rimer, L. Sperry, and J. Belnap. 2001. Exotic plant

brate fauna from Canadas High Arctic. Geol. Soc. Am. Bull.

invasion alters nitrogen dynamics in an arid grassland. Ecol.

124:323.
Eckert, Jr., R. E., F. F. Peterson, M. S. Meurisse, and J. L. Stephens.
1986. Effects of soil-surface morphology on emergence and survival of seedlings in big sagebrush communities. J. Range Manage. 39:41420.

Appl. 11:130110.
Everett, R. L., compiler. 1987. Proceedings: Pinyon-juniper conference. U.S. Forest Ser. Gen. Tech. Rep. INT-215.
Everitt, B. L. 1968. Use of the cottonwood in an investigation of the
recent history of a floodplain. Am. J. Sci. 266:41739.

Edburg, S. L., J. A. Hicke, P. D. Brooks, et al. 2012. Cascading im-

Eversman, S. T. 1968. A comparison of plant communities and sub-

pacts of bark beetlecaused tree mortality on coupled biogeo-

strates of avalanche and non-avalanche areas in south-central

physical and biogeochemical processes. Front. Ecol. Environ.


10:41624.
Eisenberg, C., S. T. Seager, and D. E. Hibbs. 2013. Wolf, elk, and
aspen food web relationships: Context and complexity. Forest
Ecol. Manage. 299:7080.
Eiseroad, A. C., and N. T. Rudd. 2011. Can Imazapic increase native species abundance in cheatgrass (Bromus tectorum) in-

Montana. M.S. thesis, Montana State Univ., Bozeman.


Eversman, S. T., and M. Carr. 1992. Yellowstone Ecology: A Road
Guide. Mountain Press Publishing, Missoula, Montana.
Ewers, B. E., and E. Pendall. 2008. Spatial patterns in leaf area and
plant functional type cover across chronosequences of sagebrush ecosystems. Plant Ecol. 194:6783.
Fahey, T. J. 1983. Nutrient dynamics of aboveground detritus in

vaded native plant communities? Rangeland Ecol. Manage.

lodgepole pine (Pinus contorta ssp. latifolia) ecosystems. Ecol.

64:64148.

Monogr. 53:5172.

Eldridge, D. J. 2004. Mounds of the American badger (Taxidea


taxus): Significant features of North American shrub-steppe
ecosystems. J. Mammalogy 85:106067.
. 2009. Badger (Taxidea taxus) mounds affect soil hydrological properties in a degraded shrub-steppe. American Midl. Naturalist 161:35058.
Elias, S. A. 2001. Paleoecology and late quaternary environments of
the Colorado Rockies. In W. D. Bowman and T. R. Seastedt, eds.,
Structure and Function of an Alpine Ecosystem, pp. 30419.
Oxford Univ. Press, New York.
Elliott, E. T., and D. C. Coleman. 1988. Let the soil work for us. Ecol.
Bull. (Sweden) 39:2332.
Elliott, G. P., and K. E. Kipfmueller. 2010. Multi-scale influences
of slope aspect and spatial pattern of ecotonal dynamics at the
upper treeline in the Southern Rocky Mountains, USA. Arctic
Antarctic Alp. Res. 42:4556.
. 2011. Multiscale influences of climate on upper treeline
dynamics in the Southern Rocky Mountains, USA: Evidence

Fahey, T. J., and D. H. Knight. 1986. Lodgepole pine ecosystems.


BioScience 36:61017.
Fahey, T. J., J. B. Yavitt, J. A. Pearson, and D. H. Knight. 1985. The
nitrogen cycle in lodgepole pine forests, southeastern Wyoming. Biogeochem. 1:25775.
Fahey, T. J., J. B. Yavitt, and G. Joyce. 1988. Precipitation and
throughfall chemistry in Pinus contorta ssp. latifolia ecosystems,
southeastern Wyoming. Can. J. Forest Res. 18:33745.
Fahnestock, J. T., and J. K. Detling. 2002. Bison-prairie dog-plant
interactions in a North American mixed-grass prairie. Oecologia 132:8695.
Fall, P. L., P. T. Davis, and G. A. Zielinski. 1995. Late quaternary
vegetation and climate of the Wind-River Range, Wyoming.
Quatern. Res. 43:393404.
Fedy, B. C., C. L. Aldredge, K. E. Doherty, et al. 2011. Interseasonal
movements of greater sage-grouse, migratory behavior, and an
assessment of the core regions concept in Wyoming. J. Wildlife
Manage. 76:106271.

of intraregional variability and bioclimatic thresholds in re-

Fenn, M. E., J. S. Baron, E. B. Allen, et al. 2003. Ecological effects

sponse to twentieth-century warming. Ann. Assoc. Am. Geogr.

of nitrogen deposition in the Western United States. BioScience

101:11811203.
Elliott-Fisk, D. L., B. S. Adkins, and J. L. Spaulding. 1983. A reevaluation of the postglacial vegetation of the Laramie Basin.
Great Basin Nat. 43:37784.
Ellison, L. 1954. Subalpine vegetation of the Wasatch Plateau, Utah.
Ecol. Monogr. 24:89184.
. 1960. Influence of grazing on plant succession of rangelands. Bot. Rev. 26:178.
Ellison, L., and E. J. Woolfolk. 1937. Effects of drought on vegetation near Miles City, Montana. Ecology 18:32936.
Engle, D. M., and P. M. Bultsma. 1984. Burning of northern mixed
prairie during drought. J. Range Manage. 37:398401.

53:40420.
Ferguson, C. W. 1964. Annual rings in big sagebrush. Paper no. l,
Laboratory of tree-ring research. Univ. Arizona Press, Tucson.
Ferreira, J. F. S., and J. Janick. 2004. Allelopathic plants. XVI. Artemisia species. Allelopathy J. 14:16775.
Fettig, C. J., K. D. Klepzig, R. F. Billings, et al. 2007. The effectiveness
of vegetation management practices for prevention and control
of bark beetle infestations in coniferous forests of the western
and southern United States. Forest Ecol. Manage. 238:2453.
Finch, D. M., ed. 2012. Climate change in grasslands, shrublands,
and deserts of the interior American West: A review and needs
assessment. U.S. Forest Ser. RMRS-GTR 285.

References
Finney, M. A. 2001. Design of regular landscape fuel treatment patterns for modifying fire growth and behavior. Forest Sci. 47:
21928.

Freilich, J. E., J. M. Emlen, J. J. Duda, et al. 2003. Ecological effects


of ranching: A six-point critique. BioScience 53:75965.
Fremont, J. C. 1845. Report of the exploring expedition to the

Finney, M. A., C. W. McHugh, R. Bartlette, et al. 2003. Fire behav-

Rocky Mountains in the year 1842, and to Oregon and North

ior, fuel treatments, and fire suppression on the Hayman Fire,

California in the years 184344. 28th Congress, Serial 461. Sen-

Part 2: Description and interpretations of fire behavior. In R. T.


Graham, ed., Hayman Fire case study, pp. 5995. U.S. Forest Ser.
Gen. Tech. Rep. RMRS GTR-114.
Fisher, R. F., M. J. Jenkins, and W. F. Fisher. 1987. Fire and the
prairie-forest mosaic of Devils Tower National Monument. Am.
Midl. Nat. 117:25057.

ate Exec. Doc. 174.


Friedman, J. M., M. L. Scott, and G. T. Auble. 1996. Water management and cottonwood forest dynamics along prairie streams. In
F. L. Knopf and F. B. Samson, eds., Ecology and Conservation of
Great Plains Vertebrates, pp. 4971. Springer, New York.
Friese, C. F. and M. F. Allen. 1993. The interaction of harvester

Flowers, S., and F. R. Evans. 1966. The flora and fauna of the Great

ants and vesicular arbuscular mycorrhizal fungi in a patchy

Salt Lake region, Utah. In H. Boyko, ed., Salinity and aridity.

semiarid environment: The effects of mound structure

Monographics Biologicae 16, pp. 36793. W. Junk, The Hague.

on fungal dispersion and establishment. Functional Ecol.

Fogel, R., and J. M. Trappe. 1978. Fungus consumption (mycophagy) by small animals. Northw. Sci. 52:131.
Forbis, T. A. 2003. Seedling demography in an alpine ecosystem.
American Journal of Botany 90:11971206.

7:1320.
Friggens, M. M., M. V. Warwell, J. C. Chambers, and S. G. Kitchen.
2012. Modeling and predicting vegetation response of western
USA grasslands, shrublands, and deserts to climate change. In

Forman, S. L., and J. Pierson. 2003. Formation of linear parabolic

D.M. Finch, ed., Climate change in grasslands, shrublands, and

dunes on the eastern Snake River Plain, Idaho in the nineteenth

deserts of the interior American West: A review and needs as-

century. Geomorphology 56:189200.


Forman, S. L., L. Marin, J. Pierson, et al. 2005. Aeolian sand depositional records from western Nebraska: Landscape response to
droughts in the past 1500 years. Holocene 15:97381.

sessment, pp. 120. U.S. Forest Serv. RMRS-GTR 285, Fort Collins, Colorado.
Frischknecht, N. C., and M. F. Baker. 1972. Voles can improve sagebrush rangelands. J. Range Manage. 25:46668.

Forrest, J., D. W. Inouye, and J. D. Thomson. 2010. Flowering phe-

Frischknecht, N. C., and A. P. Plummer. 1955. A comparison of

nology in subalpine meadows: Does climate variation influence

seeded grasses under grazing and protection on a mountain

community co-flowering patterns? Ecology 91:43140.


Fortin, J. K., C. C. Schwartz, K. A. Gunther, et al. 2013. Dietary adjustability of grizzly bears and American black bears in Yellow
stone National Park. J. Wildlife Manage. 77:27081.
Foster, M. A., and J. Stubbendieck. 1980. Effects of the plains
pocket gopher (Geomys bursarius) on rangeland. J. Range Manage. 33:7478.
Foster, R. H. 1968. Distribution of the major plant communities in
Utah. Ph.D. diss., Brigham Young Univ., Provo, Utah.
Frank, D. A., and R. D. Evans. 1997. Effects of native grazers on
grassland N cycling in Yellowstone National Park. Ecology
78:223848.
Frank, D. A., and S. J. McNaughton. 1992. The ecology of plants, large
mammalian herbivores, and drought in Yellowstone National
Park. Ecology 73:204358.
. 1993. Plant-ungulate ecology of grasslands on the northern range of Yellowstone National Park. In D. G. Despain

brush burn. J. Range Manage. 8:17075.


Frison, G. C. 1991. Prehistoric Hunters of the High Plains, 2nd edition. Academic Press, New York.
Froiland, S. G. 1990. Natural History of the Black Hills and Badlands. Center for Western Studies, Augustana College, Sioux
Falls, South Dakota.
Frost, S. M., P. D. Stahl, and S. E. Williams. 2001. Long-term reestablishment of arbuscular mycorrhizal fungi in a drastically
disturbed semiarid surface mine soil. Arid Land Res. and Manage. 15:312.
Fryxell, F. 1978. Mountaineering in the Tetons: The Pioneer Period 18981940. P. D. Smith, ed. Teton Bookshop, Jackson,
Wyoming.
Fuhlendorf, S. D., D. M. Engle, R. D. Elmore, et al. 2012. Conservation of pattern and process: Developing an alternative paradigm of rangeland management. Rangeland Ecol. Manage. 65:
57989.

and R. H. Hamre, eds. Plants and their environments: First

Furniss, M. M., and R. G. Krebill. 1972. Insects and diseases of

Bi
ennial Scientific Conference on the Greater Yellowstone

shrubs on western big game ranges. In D. M. McKell, J. P. Blais-

Ecosystem. National Park Serv. Tech. Rep., U.S. Gov. Printing

dell, and J. R. Goodin, eds., Wildland shrubsTheir biology

Office, Denver.

and utilization, pp. 21826. U.S. Forest Ser. Gen. Tech. Rep.

Frank, D. A., P. M. Groffman, R. D. Evans, and B. F. Tracy. 2000. Ungulate stimulation of nitrogen cycling and retention in Yellow
stone Park grasslands. Oecologia 123:11621.
Frank, D. A., T. Depriest, K. McLauchlan, and A. D. Risch. 2011. Topo
graphic and ungulate regulation of soil C turnover in a temperate
grassland ecosystem. Global Change Biol. 17:495504.
Freeland, R. O. 1944. Apparent photosynthesis in some conifers
during winter. Plant Phys. 19:17985.

INT-1.
Gabet, E., J. Perron, J. Taylor, et al. 2014. Biotic origin for mima
mounds supported by numerical modeling. Geomorphology
206:5866.
Gage, E., and D. J. Cooper. 2012. Historical Range of Variation Assessments for Wetland and Riparian Ecosystems, U.S. Forest
Service Rocky Mountain Region. Dept. Forest and Rangeland
Stewardship, Colorado State Univ., Fort Collins.

359

360 References
Gammonley, J. H. 2004. Wildlife of natural palustrine wetlands.

Goldblum, D., and T. T. Veblen. 1992. Fire history of a ponderosa

In M. C. McKinstry, W. A. Hubert, and S. H. Anderson, eds.,

pine
/
Douglas-fir forest in the Colorado front range. Phys.

Wetland and Riparian Areas of the Intermountain West: Ecol-

Geogr. 13:13348.

ogy and Management, pp. 13053. Univ. Texas Press, Austin.


Ganskopp, D. C. 1986. Tolerances of sagebrush, rabbitbrush, and
greasewood to elevated water tables. J. Range Manage. 39:33437.
Gardner, D. 2008a. Environmental change in the Wyoming Basins
Red Desert. In A. Proulx, ed., Red Desert: History of a Place, pp.
12135. Univ. Texas Press, Austin.
. 2008b. Horses come to the Red Desert. In A. Proulx, ed., Red
Desert: History of a Place, pp. 18187. Univ. Texas Press, Austin.

Good, J. M., and K. L. Pierce. 2002. Interpreting the landscapes


of Grand Teton and Yellowstone National Parks: Recent and
ongoing geology. Grand Teton Nat. Hist. Assoc., Moose,
Wyoming.
Goodin, J. R., and A. Mozafar. 1972. Physiology of salinity stress.
In C. M. McKell, J. P. Blaisdell, and J. R. Goodin, eds., Wildland
Shrubs: Their Biology and Utilization, pp. 25559, U.S. Forest
Ser. Gen. Tech. Rep. INT-1.

Gary, H. L., and C. A. Troendle. 1982. Snow accumulation and melt

Gorham, E., P. M. Vitousek, and W. A. Reiners. 1979. The regulation

under various stand densities in lodgepole pine in Wyoming

of chemical budgets over the course of terrestrial ecosystem suc-

and Colorado. U.S. Forest Ser. Res. Note RM-417.


Gates, D. H., L. A. Stoddart, and C. W. Cook. 1956. Soil as a factor
in influencing plant distribution on salt deserts of Utah. Ecol.
Monogr. 26:15575.
Gavin, D. G., D. J. Hallett, F. S. Hu, et al. 2007. Forest fire and climate change in western North America: Insights from sediment
charcoal records. Front. Ecol. Environ. 5:499506.
Gaylord, D. R. 1982. Geologic history of the Ferris dune field,

cession. Annu. Rev. Ecol. Syst. 10:5384.


Gosz, J. R. 1980. Biomass distribution and production budget for a
nonaggrading forest ecosystem. Ecology 61:50714.
Grafe, E., and P. Horsted. 2002. Exploring with Custer: The 1874
Black Hills Expedition. Golden Valley Press, Custer, South
Dakota.
Grande, L. 1998. Locked in stone: An extinct 50-million-year-old
lake system. Nat. Hist. 107:6669.

south-central Wyoming. In R. W. Marrs and K. E. Kolm, eds.,

Grasso, D. 1990. Recognition and paleogeography of Quaternary

Interpretation of Windflow Characteristics from Eolian Land-

Lake Wamsutter (a proposed lake in Wyomings Great Divide

forms, pp. 6582. Special Paper 192. Geol. Soc. Am., Boulder.

Basin) combining Landsat remote sensing and digital elevation

Geils, B. W., and W. R. Jacobi. 1990. Development of comandra blister rust on lodgepole pine. Can. J. Forest Res. 20:15965.

modeling. Ph.D. diss., Univ. Wyoming, Laramie.


Gray, S. T., and C. Andersen. 2009. Assessing the Future of Wyo-

Gemmill, C. E. C., and K. R. Johnson. 1997. Paleoecology of a late

mings Water Resources: Adding Climate Change to the Equa-

Paleocene (Tiffian) megaflora from the northern Great Divide

tion. William D. Ruckelshaus Institute of Environment and

Basin, Wyoming. Palaios 12:43948.

Natural Resources, Univ. Wyoming, Laramie. www.uwyo.edu/

Gibbens, R. P. 1972. Vegetation pattern within northern desert


shrub communities. Ph.D. diss., Univ. Wyoming, Laramie.
Gifford, G. F., and R. H. Hawkins. 1978. Hydrologic impact of
grazing on infiltration: A critical review. Water Resour. Res.
14:30513.
Gilbert, M. M., and A. D. Chalfoun. 2011. Energy development affects populations of sagebrush songbirds in Wyoming. J. Wildlife Manage. 75:81624.
Gill, R. A., and I. C. Burke. 1999. Ecosystem consequences of plant
life form changes at three sites in the semiarid United States.
Oecologia 121:55163.

enr.
Gray, S. T., C. L. Fastie, S. T. Jackson, and J. L. Betancourt. 2004.
Tree-ring based reconstructions of precipitation in the Bighorn
Basin, Wyoming since A.D. 1260. J. Climate 17:385565.
Gray, S. T., J. L. Betancourt, S. T. Jackson, and R. G. Eddy. 2006.
Role of multidecadal climate variability in a range extension of
pinyon pine. Ecology 87:112430.
Grayson, D. K. 2006. Holocene bison in the Great Basin, western
USA. Holocene 6:91325.
Green, A. W., and R. C. Conner. 1989. Forests in Wyoming. U.S.
Forest Ser. Res. Bull. INT-61.

Gingerich, P. D., and W. C. Clyde. 2001. Overview of mammalian

Gresswell, R. E., and W. J. Liss. 1995. Values associated with man-

biostratigraphy in the Paleocene-Eocene Fort Union and Will-

agement of Yellowstone cutthroat trout in Yellowstone National

wood formations of the Bighorn and Clarks Fork basins. In

Park. Conserv. Biol. 9:15965.

P. D. Gingerich, ed., Paleocene-Ecocene Stratigraphy and Bi-

Gribb, W. J., and D. J. Brosz. 1990. Irrigated acreage. In L. M.

otic Change in the Bighorn and Clarks Fork Basins, Wyoming,

Ostresh, Jr., R. A. Marston, and W. M. Hudson, eds., Wyoming

pp.114. Papers on Paleontology 33. Univ. Michigan, Ann Arbor.


Girard, M. M., D. L. Wheeler, and S. B. Mills. 1997. Classification of
riparian communities on the Bighorn National Forest. U.S. Forest Ser. Bighorn National Forest, Sheridan, Wyoming.
Glenn, E. P., K. Morino, P. L. Nagler, et al. 2012. Roles of saltcedar
(Tamarix spp.) and capillary rise in salinizing a non-flooding terrace on a flow-regulated desert river. J. Arid Environ. 79:5665.

Water Atlas, pp. 3839. Wyoming Water Development Commission, Cheyenne.


Grier, C. C., and S. W. Running. 1977. Leaf area of mature northwestern coniferous forests: Relation to site water balance. Ecology 58:89399.
Gries, J. P. 1996. Roadside Geology of South Dakota. Mountain
Press, Missoula, Montana.

Goergen, E. M., E. A. Leger, and E. K. Espeland. 2011. Native peren

Griffin, J. M., and M .G. Turner. 2012. Changes to the N cycle fol-

nial grasses show evolutionary response to Bromus tectorum

lowing bark beetle outbreaks in two contrasting conifer forest

(cheatgrass) invasion. PLoS ONE. 6:e18145.

types. Oecologia 170:55165.

References
Griffin, J. M., M. G. Turner, and M. Simard. 2011. Nitrogen cycling

Hansen, A. J., R. Rasker, B. Maxwell, et al. 2002. Ecological causes

following mountain pine beetle disturbance in lodgepole pine

and consequences of demographic change in the New West.

forests of Greater Yellowstone. Forest Ecol. Manage. 261:107789.

BioScience 52:15162.

Griggs, R. F. 1946. The timberlines of North America and their interpretation. Ecology 27:27589.
Gruell, G. E. 1983. Fire and vegetative trends in the northern
Rockies: Interpretations from 18711982 photographs. U.S. Forest Ser. Gen. Tech. Rep. INT-158.
Gruell, G. E., and L. L. Loope. 1974. Relationships among aspen,

Hansen, K., A. Jakle, and M. Hogarty. 2013. Market-based wildlife


mitigation in Wyoming: A primer. Ruckelshaus Institute of
Environment and Natural Resources, Laramie, Wyoming.
Hansen, P. L. 1995. Classification and management of Montanas
riparian and wetland sites. Bull. 54. Montana Forest Conserv.
Exp. Sta., School of Forestry, Univ. Montana, Missoula.

fire, and ungulate browsing in Jackson Hole, Wyoming. U.S.

Hansen, P. L., and G. R. Hoffman. 1988. The vegetation of the

Forest Ser. (Intermountain Region) and U.S. National Park Ser-

Grand River/Cedar River, Sioux, and Ashland Districts of the

vice. http://digitalcommons.usu.edu/aspen_bib/5239.

Custer National Forest: A habitat type classification. U.S. Forest

Gruell, G. E., S. Bunting, and L. Neuenschwander. 1985. Influence


of fire on curlleaf mountain-mahogany in the intermountain
west. In J. K. Brown and J. Lotan, eds., Fires effects on wildlife
habitat, pp. 5872. U.S. Forest Ser. Gen. Tech. Rep. INT-186.
Gude, P. H., A. J. Hansen, and D. A. Jones. 2007. Biodiversity con-

Ser. Gen. Tech. Rep. RM-157.


Hansen, R. M. 1962. Movements and survival of Thomomys talpo
ides in a mima-mound habitat. Ecology 43:15154.
Hanson, C. L., C. W. Johnson, and J. R. Wight. 1982. Foliage mortality of mountain big sagebrush (Artemisia tridentata ssp.

sequences of alternative future land use scenarios in Greater

vaseyana) in southwestern Idaho during the winter of 197677.

Yellowstone. Ecol. Appl. 17:10048.

J. Range Manage. 35:14245.

Gullap, M. K., H. I. Erkovan, and A. Koc. 2011. The effect of bovine

Hanson, P. R., R. M. Joeckel, A. R. Young, and J. Horn. 2009. Late

saliva on growth attributes and forage quality of two contrast-

Holocene dune activity in the Eastern Platte River Valley,

ing cool season grasses grown in three soils of different fertility.


Rangeland J. 33:30713.
Gurevitch, J., S. M. Scheiner, and G. A. Fox. 2002. The Ecology of
Plants. Sinauer Associates, Sunderland, Massachusetts.
Hadley, J. L., and W. K. Smith. 1986. Wind effects on needles of
timberline conifers: Seasonal influence on mortality. Ecology
67:1219.
. 1987. Influence of krummholz mat structure on micro
climate and needle physiology. Oecologia 73:8290.
. 1989. Wind erosion of leaf surface wax in alpine timberline
conifers. Arct. Alp. Res. 21:39298.
Haines, A. L., ed. 1965. Osborne Russells Journal of a Trapper. Univ.
Nebraska Press, Lincoln.
. 1996. The Yellowstone Story: A History of Our First National
Park, revised edition, vol. 2. Univ. Press Colorado, Boulder.
Halfen, A. F., G. G. Fredund, and S. A. Mahan. 2010. Holocene
stratigraphy and chronology of the Casper Dune Field, Casper,
Wyoming, USA. Holocene 20:77383.
Hamerlinck, J. D., S. N. Lieske, and W. J. Gribb. 2013. Understanding Wyoming Land Resources: Land-use Patterns and Development Trends. Ext. Publ. B-1244. Univ. Wyoming, Laramie.
Hamerlynck, E. 1992. Subnivian and emergent growth, morphol-

Nebraska. Geomorphology 103:55561.


Hanson, P. R., A. F. Arbogast, W. C. Johnson, et al. 2010. Megadroughts and late Holocene dune activation at the eastern margin of the Great Plains, north-central Kansas, USA. Aeolian Res.
1:10110.
Harder, L. D. 1979. Winter feeding by porcupines in southwestern
Alberta: Preferences, density effects, and temporal changes.
Can. J. Zool. 58:1319.
Harmon, M. E., J. F. Franklin, F. J. Swanson, et al. 1986. Ecology of
coarse woody debris in temperate ecosystems. Adv. Ecol. Res.
15:133302.
Harniss, R. O., and R. B. Murray. 1973. Thirty years of vegetal
change following burning of sagebrush-grass range. J. Range
Manage. 26:32225.
Harper, K. T., F. J. Wagstaff, and L. M. Kunzler. 1985. Biology and
management of the Gambel oak vegetative type: A literature
review. U.S. Forest Ser. Gen. Tech. Rep. INT-179.
Harr, R. D., and K. R. Price. 1972. Evapotranspiration from
a greasewood-cheatgrass community. Water Resour. Res.
8:11991203.
Harris, T. 1992. David Love: Warnings about selenium. High Country News, Feb. 10, pp. 1, 812.

ogy, and physiology of Erythronium grandiflorum in relation to

Hart, E. A., and J. R. Lovvorn. 2000. Vegetation dynamics and

snow melt dynamics in the subalpine forest, Medicine Bow

primary production in saline, lacustrine wetlands of a Rocky

Mountains, Wyoming, U.S.A. M.S. thesis, Univ. Wyoming,


Laramie.
Hamner, R. W. 1964. An ecological study of Sarcobatus vermiculatus communities of the Big Horn Basin, Wyoming. M.S. thesis,
Univ. Wyoming, Laramie.
Hansen, A. J., and R. DeFries. 2007. Ecological mechanisms linking protected areas to surrounding lands. Ecol. Appl. 17:97488.

Mountain basin. Aquat. Bot. 66:2139.


Hart, R. 2001. Where the buffalo roamedOr did they? Great Plains
Res.

paper

543.

http://digitalcommons.unl.edu/greatplains

research/543.
Hart, R. H. 2008. Land-use history on the shortgrass steppe. In
W.K. Lauenroth and I. C. Burke, eds., Ecology of the Shortgrass
Steppe, pp. 5569. Oxford Univ. Press, Oxford.

Hansen, A. J., and J. J. Rotella. 2002. Biophysical factors, land use,

Hart, R. H., and W. A. Laycock. 1996. Repeat photography on range

and species viability in and around nature reserves. Conserv.

and forest lands in the western United States. J. Range Manage.

Biol. 16:111222.

49:6067.

361

362 References
Harte, J., and R. Shaw. 1995. Shifting dominance within a montane

Hiemstra, C. A., G. E. Liston, and W. A. Reiners. 2006. Observing,

vegetation community: Results of a climate-warming experi-

modeling, and validating snow redistribution by wind in a Wy-

ment. Science 267:87680.

oming upper treeline landscape. Ecol. Model. 197:3551.

Harvey, A. E., M. F. Jurgensen, M. J. Larsen, and R. T. Graham. 1987.

Higgins, K. F. 1984. Lightning fires in North Dakota grasslands and

Relationships among soil microsite, ectomycorrhizae, and natu-

in pine-savanna lands of South Dakota and Montana. J. Range

ral conifer regeneration of old-growth forests in western Montana. Can. J. Forest Res. 17:5862.
Hassan, M. A., and N. E. West. 1986. Dynamics of soil seed pools

Manage. 37:100103.
Higuera, P. E., C. Whitlock, and J. A. Gage. 2011. Linking tree-ring
and sediment-charcoal records to reconstruct fire occurrence

in burned and unburned sagebrush semideserts. Ecology 67:

and area burned in subalpine forests of Yellowstone National

26972.

Park, USA. Holocene 21:32741.

Httenschwiler, S., and W. K. Smith. 1999. Seedling occurrence in

Hild, A. L, G. E. Schuman, L. E. Vicklund, and M. I. Williams. 2006.

alpine treeline conifers: A case study from the central Rocky

Canopy growth and density of Wyoming big sagebrush sown

Mountains. Acta Oecol. 20:21924.

with cool-season perennial grasses. Arid Land Res. Manage. 20:

Havstad, K. M., D. P. C. Peters, R. Skaggs, et al. 2007. Ecological


services to and from rangelands of the United States. Ecol. Econ.
64:26168.
Hawksworth, F. G., and D. W. Johnson. 1989. Biology and management of dwarf mistletoe in lodgepole pine in the Rocky Mountains. U.S. Forest Ser. Gen. Tech. Rep. RM-169.

18394.
Hodgkinson, H. S. 1987. Relationship of saltbush species to soil
chemical properties. J. Range Manage. 40:2326.
Hoff, R. J. 1992. How to recognize blister rust infection on whitebark pine. U.S. Forest Ser. Res. Note INT-406.
Hoffman, G. R., and R. R. Alexander. 1987. Forest vegetation of

Hayden, F. V. 1879. Eleventh annual report of the United States

the Black Hills National Forest of South Dakota and Wyo-

geological and geographical survey of the territories, embracing

ming: A habitat type classification. U.S. Forest Ser. Res. Paper

Idaho and Wyoming, being a report of progress in the exploration for the year 1877. U.S. Gov. Printing Office, Washington,
D.C.
Hayes, M. 2012. Alterations to riparian vegetation resulting from
beaver pond life histories. M.S. thesis, Univ. Wyoming, Laramie.
Hebblewhite, M., C. A. White, C. G. Nietvelt, et al. 2005. Human
activity mediates a trophic cascade caused by wolves. Ecology
86:213544.
Heitschmidt, R. K. 1990. The role of livestock and other herbivores
in improving rangeland vegetation. Rangelands 12:11215.
Hennon, P. E., D. V. DAmore, P. G. Schaberg, et al. 2012. Shifting
climate, altered niche, and a dynamic conservation strategy for
yellow-cedar in the north Pacific coastal rainforest. BioScience
62:14758.
Hess, J. E., and J. L. Beck. 2012a. Burning and mowing Wyoming

RM-276.
Hoffman, G. R., and D. L. Hazlett. 1977. Effects of aqueous Artemisia extracts and volatile substances on germination of selected species. J. Range Manage. 30:13437.
Holland, E. A., and J. K. Detling. 1990. Plant response to herbivory
and belowground nitrogen cycling. Ecology 71:104049.
Holloran, M. J., B. J. Heath, A. G. Lyon, et al. 2005. Greater sagegrouse nesting habitat selection and success in Wyoming.
J.Wildlife Manage. 69:63849.
Holmes, A. 2007. Short-term effects of a prescribed burn on songbirds and vegetation in mountain big sagebrush. West. N. Am.
Nat. 67:29298.
Holmes, N. D., D. S. Smith, and A. Johnston. 1979. Effect of grazing
by cattle on the abundance of grasshoppers on fescue grassland.
J. Range Manage. 32:31011.

big sagebrush: Do treated sites meet minimum guidelines for

Holpp, F. A. 1977. Vegetative composition and soil analysis of se-

greater sage-grouse breeding habitats? Wildlife Soc. Bull. 36:

lected playas of Campbell County, Wyoming. M.S. thesis, Univ.

8593.

Wyoming, Laramie.

. 2012b. Disturbance factors influencing greater sage-grouse

Holtmeier, F. K., and G. Broll. 1992. The influence of tree islands

lek abandonment in north-central Wyoming. J. Wildlife Man-

and microtopography on pedoecological condition in the forest

age. 76:162534.

alpine tundra ecotone on Niwot Ridge, Colorado Front Range,

Hessl, A. E., and L. J. Graumlich. 2002. Interactive effects of


human activities, herbivory and fire on quaking aspen (Populus tremuloides) age structures in western Wyoming. J. Biogeogr.
29:889902.
Heuer, K., P. D. Brooks, and K. A. Tonnessen. 1999. Nitrogen dy-

USA. Arct. Alp. Res. 24:21628.


Homsher, L. M., ed. 1960. South Pass, 1868: James Chisholms Journal of the Wyoming Gold Rush. Univ. Nebraska Press, Lincoln.
. 1965. The History of Albany County, Wyoming. Lusk Herald, Lusk, Wyoming.

namics in two high elevation catchments during spring snow-

Hood, E. W., M. W. Williams, and N. Caine. 2003. Landscape con-

melt 1996, Rocky Mountains, Colorado. Hydrol. Process. 13:

trols on organic and inorganic nitrogen leaching across and

220314.

alpine/subalpine ecotone, Green Lakes Valley, Colorado Front

Hewitt, G. B., and J. A. Onsager. 1983. Control of grasshoppers on


rangeland in the United States. J. Range Manage. 36:2027.
Hicke, J. A., M. C. Johnson, J. L. Haynes, and H. K. Preisler. 2012.
Effects of bark beetlecaused tree mortality on wildfire. Forest
Ecol. Manage. 271:8190.

Range. Ecosystems 6:3145.


Houston, D. B. 1967. The Shiras moose in Jackson Hole, Wyoming.
Ph.D. diss., Univ. Wyoming, Laramie.
. 1973. Wildfires in northern Yellowstone National Park.
Ecology 54:111117.

References
. 1982. The Northern Yellowstone Elk: Ecology and Management. MacMillan, New York.
Houston, W. R. 1961. Some interrelations of sagebrush, soils, and
grazing intensity in the northern Great Plains. Ecology 42:
3138.
. 1963. Plains pricklypear, weather, and grazing in the northern Great Plains. Ecology 44:56974.
Houston, W. R., and D. N. Hyder. 1975. Ecological effect and fate
of N following massive N fertilization of mixed grass plains.
J.Range Manage. 28:5660.
Houston, W. R., L. D. Sabatka, and D. N. Hyder. 1973. Nitratenitrogen accumulation in range plants after massive N fertilization on shortgrass plains. J. Range Manage. 26:5457.

Inouye, D. W. 2008. Effects of climate change on phenology, frost


damage, and floral abundance of montane wildflowers. Ecology
89:35362.
Inouye, R. S. 2006. Effects of shrub removal and nitrogen addition on soil moisture in sagebrush steppe. J. Arid Environ. 66:
60418.
Irvine, J. R., and N. E. West. 1979. Riparian tree species distribution
and succession along the lower Escalante River, Utah. Southw.
Nat. 24:33146.
Jackson, S. T., M. E. Lyford, and J. L. Betancourt. 2002. A 4000-year
record of woodland vegetation from Wind River Canyon, central Wyoming. West. N. Am. Nat. 62:40513.
Jackson, S. T., J. L. Betancourt, M. E. Lyford, et al. 2005. A 40,000-

Hubert, W. A. 2004. Ecological processes of riverine wetland habi-

year woodrat-midden record of vegetational and biogeographi-

tats. In M. C. McKinstry, W. A. Hubert, and S. H. Anderson,

cal dynamics in north-eastern Utah, USA. J. Biogeogr. 32:

eds., Wetland and Riparian Areas of the Intermountain West,


pp. 5273. Univ. Texas Press, Austin.
Hudson, H. J. 1986. Fungal Biology. Edward Arnold, London.
Hultine, K. R., and S. E. Bush. 2011. Ecohydrological consequences

10851106.
Jackson, W. N. 1957. Some soil characteristics of several grasslandtimber transitions in the Bighorn Mountains and the Laramie
Plains. M.S. thesis, Univ. Wyoming, Laramie.

of non-native riparian vegetation in the southwestern United

Jacobi, W. R., B. W. Geils, J. E. Taylor, and W. R. Zentz. 1993. Pre-

States: A review from an ecophysiological perspective. Water

dicting the incidence of comandra blister rust on lodgepole

Resour. Res. 47:W07542.

pine: Site, stand, and alternate host influences. Phytopathology

Hultine, K. R., J. Belnap, C. van Riper III, et al. 2010a. Tamarisk


biocontrol in the western United States: Ecological and societal
implications. Front. Ecol. Environ. 8:46774.
Hultine, K. R., S. E. Bush, and J. R. Ehleringer. 2010b. Ecophysiology
of riparian cottonwood and willow before, during, and after
two years of soil water removal. Ecol. Appl. 20:34761.
Hunt, H. W., D. C. Coleman, E. R. Ingham, et al. 1987. The detrital
food web in a shortgrass prairie. Biol. Fertil. Soils 3:5768.
Huntly, N. J. 1987. Influence of refuging consumers (pikas: Ochotona
princeps) on subalpine meadow vegetation. Ecology 68:27483.
Huntly, N. J., and R. Inouye. 1988. Pocket gophers in ecosystems:
Patterns and mechanisms. BioScience 38:78693.
Hurd, R. M. 1961. Grassland vegetation in the Big Horn Mountains,
Wyoming. Ecology 42:45967.
Hutchins, H. E., and R. M. Lanner. 1982. The central role of Clarks
nutcracker in the dispersal and establishment of whitebark
pine. Oecologia 55:19220l.
Hutto, R. L. 2008. The ecological importance of severe wildfires:
Some like it hot. Ecol. Appl. 18:182734.
Huxman, T. E., K. A. Snyder, D. Tissue, et al. 2004. Precipitation
pulses and carbon fluxes in semiarid and arid ecosystems.
Oecologia 141:25468.
Ingelfinger, F., and S. Anderson. 2004. Passerine response to roads
associated with natural gas extraction in a sagebrush steppe
habitat. West. N. Am. Nat. 64:38595.

83:63037.
Jakle, A. 2012a. Natural Gas Development and Wildlife Mitigation:
A Primer. Ruckelshaus Institute of Environment and Natural
Resources, Laramie, Wyoming.
. 2012b. Wind Development and Wildlife Mitigation in
Wyoming: A Primer. Ruckelshaus Institute of Environment and
Natural Resources, Laramie, Wyoming.
Jakubos, B., and W. H. Romme. 1993. Invasion of subalpine meadows by lodgepole pine in Yellowstone National Park, Wyoming,
USA. Arct. Alp. Res. 25:38290.
James, J. J., R. L. Tiller, and J. H. Richards. 2005. Multiple resources
limit plant growth and function in a saline-alkaline desert
community. J. Ecol. 93:11326.
Janetski, J. C. 1987. Indians of Yellowstone National Park. Univ.
Utah Press, Salt Lake City.
Jansen, E., J. Overpeck, K. R. Briffa, et al. 2007. Palaeoclimate. In
S. Solomon, D. Qin, M. Manning, eds., Climate Change: The
Physical Science Basis. Contribution of Working Group I to the
Fourth Assessment Report of the Intergovernmental Panel on
Climate Change. Cambridge Univ. Press, Cambridge.
Jaramillo, V. J., and J. K. Detling. 1988. Grazing history, defoliation,
and competition: Effects on shortgrass production and nitrogen accumulation. Ecology 69:15991608.
Jarchow, M. E., and B. J. Cook. 2009. Allelopathy as a mechanism
for the invasion of Typha angustifolia. Plant Ecol. 204:11324.

Ingham, R. E., and J. K. Detling. 1984. Plant-herbivore interactions

Jaynes, R. A. 1978. A hydrologic model of aspen-conifer succes-

in a North American mixed-grass prairie 3: Soil nematode pop-

sion in the western United States. U.S. Forest Ser. Res. Paper

ulations and root biomass on Cynomys ludovicianus colonies and


adjacent uncolonized areas. Oecologia 63:30713.
Ingham, R. E., J. A. Trofymow, E. R. Ingham, and D. C. Coleman.

INT-213.
Jefferies, R. 1981. Osmotic adjustment and the response of halophytic plants to salinity. BioScience 31:4246.

1985. Interactions of bacteria, fungi, and their nematode graz-

Johnson, B. G., D. W. Johnson, J. C. Chambers, and R. R. Blank.

ers: Effects on nutrient cycling and plant growth. Ecol. Monogr.

2011. Fire effects on the mobilization and uptake of nitrogen by

55:11940.

cheatgrass. Plant Soil 341:43745.

363

364 References
Johnson, D. A., and M. D. Rumbaugh. 1981. Nodulation and acety-

Kauffman, J. B., and W. C. Krueger. 1984. Livestock impacts on

lene reduction by certain rangeland legume species under field

riparian ecosystems and streamside management implications:

conditions. J. Range Manage. 34:78181.

A review. J. Range Manage. 37:43038.

Johnson, D. F. 1950. Plant succession on the Missouri River flood-

Kauffman, J. B., W. C. Krueger, and M. Vavra. 1983. Impacts of cat-

plain near Vermillion, South Dakota. M.A. thesis, Univ. South

tle on streambanks in northeastern Oregon. J. Range Manage.

Dakota, Vermillion.

36:68385.

Johnson, E. A. 1987. The relative importance of snow avalanche

Kauffman, M. J., J. F. Brodie, and E. S. Jules. 2010. Are wolves sav-

disturbance and thinning on canopy plant populations. Ecol-

ing Yellowstones aspen? A landscape-level test of a behaviorally

ogy 68:4353.
Johnson, E. A., and G. I. Fryer. 1989. Population dynamics in lodgepole pineEngelmann spruce forests. Ecology 70:133545.
Johnson, E. A., and S. L. Gutsell. 1993. Heat budget and fire behavior associated with the opening of serotinous cones in two Pinus
species. J. Veg. Sci. 4:74550.
Johnson, K. L. 1987. Rangeland through time: A photographic
study of vegetation change in Wyoming 18701986. Misc. Publ.
50. Wyoming Agric. Exp. Sta., Laramie.
Johnson, K. R., and R. G. Raynolds. 2003. Ancient Denvers. Denver
Museum of Nature and Science, Denver.
Johnson, P. L., and W. D. Billings. 1962. The alpine vegetation of
the Beartooth Plateau in relation to cryopedogenic processes
and patterns. Ecol. Monogr. 32:10535.

mediated trophic cascade. Ecology 91:274255.


Kaufmann, M. R. 1985. Annual transpiration in subalpine forests:
Large differences among four tree species. Forest Ecol. Manage.
13:23546.
Kaufmann, M. R., and R. K. Watkins. 1990. Characteristics of highand low-vigor lodgepole pine trees in old-growth stands. Tree
Physiol. 7:23946.
Kaufmann, M. R., C. A. Troendle, M. G. Ryan, and H. Todd Mowrer. 1987. Trees: The link between silviculture and hydrology. In
C.A. Troendle, M. R. Kaufmann, R. H. Hamre, and R. P. Wino
kur, tech. coords., Management of subalpine forests: Building
on fifty years of research, pp. 5460. U.S. Forest Ser. Gen. Tech.
Rep. RM-149.
Kay, C. E. 1993. Historical conditions of woody vegetation on

Johnson, W. C., W. B. Werner, G. R. Guntenspergen, et al. 2010.

Yellowstones Northern Range: A critical evaluation of the nat-

Prairie wetland complexes as landscape functional units in a

ural regulation paradigm. In D. G. Despain and R. H. Hamre,

changing climate. BioScience 60:12840.

eds., Plants and their environments: First biennial scientific

Johnson, W. M. 1969. Life expectancy of a sagebrush control in


central Wyoming. J. Range Manage. 22:17782.

conference on the Greater Yellowstone Ecosystem. National


Park Serv. Tech. Rep., U.S. Gov. Printing Office, Denver.

Johnson-Nistler, C. M., B. F. Sowell, H. W. Sherwood, and C. L.

Kayes, L. J., and D. B. Tinker. 2012. Forest structure and regenera-

Wambolt. 2004. Black-tailed prairie dog effects on Montanas

tion following a mountain pine beetle epidemic in southeast-

mixed-grass prairie. J. Range Manage. 57:64148.

ern Wyoming. Forest Ecol. Manage. 263:5766.

Joyce, L. A. 1981. Climate/vegetation relationships in the northern

Keane, R. E., and A. W. Schoettle. 2011. Strategies, tools, and chal-

Great Plains and the Wyoming northcentral basins. Ph.D. diss.,

lenges for sustaining and restoring high elevation five-needle

Colorado State Univ., Fort Collins.

white pine forests in western North America. In R. E. Keane,

Kamm, J. A., F. A. Sneva, and L. M. Rittenhouse. 1978. Insect graz-

D. F. Tomback, M. P. Murray, and C. M. Smith, eds., The future

ers on the cold desert biome. In D. N. Hyder, ed., First Interna-

of high-elevation, five-needle white pines in Western North

tional Rangeland Congress, pp. 47983. Soc. Range Manage.,

America: Proceedings of the High Five Symposium: The Future

Denver.

of High-elevation, Five-needle White Pines in Western North

Karban, R., K. Shiojiri, and S. Ishizaki. 2011. Plant communication:


Why should plants emit volatile cues? J. Plant Interact. 6:8184.
Kashian, D. M., M. G. Turner, W. H. Romme, and C. G. Lorimer.

America, pp. 27694. U.S. Forest Serv. RMRS-P-63.


Keddy, P. A. 2010. Wetland Ecology: Principles and Conservation,
2nd edition. Cambridge Univ. Press, Cambridge.

2005. Variability and convergence in stand structural devel-

Kelly, E. F., C. M. Yonker, S. W. Blecker, and C. G. Olson. 2008.

opment on a fire-dominated subalpine landscape. Ecology

Soil development and distribution in the shortgrass steppe eco

86:64354.

system. In W. K. Lauenroth and I. C. Burke, eds., Ecology of the

Kashian, D. M., W. H. Romme, and C. M. Regan. 2007. Variation in


quaking aspen dynamics and decline in the northern Colorado
Front Range, U.S.A. Ecol. Appl. 17:12961311.
Kashian, D. M., W. H. Romme, D. B. Tinker, et al. 2006. Carbon
storage on landscpes with stand-replacing fires. BioScience
56:598606.
. 2013. Postfire changes in forest carbon storage over a 300year chronosequence of Pinus contortadominated forests. Ecol.
Monogr. 83:4966.

Shortgrass Steppe, pp. 3054. Oxford Univ. Press, Oxford.


Khan, M. A., B. Gul, and D. J. Weber. 2002. Seed germination in
relation to salinity and temperature in Sarcobatus vermiculatus.
Biol. Plant. 45:13335.
Kiesecker, J. M., T. Comendant, T. Grandmason, et al. 2007. Conservation easements in context: A quantitative analysis of
their use by The Nature Conservancy. Front. Ecol. Environ.
5:12530.
Kiesecker, J. M., H. Copeland, A. Pocewicz, and B. McKenney.

Katz, G. L., and P. B. Shafroth. 2003. Biology, ecology and man-

2009a. Development by design: Blending landscape-level

agement of Elaeagnus angustifolia L. (Russian olive) in western

planning with the mitigation hierarchy. Front. Ecol. Environ.

North America. Wetlands 23:76377.

8:26166.

References
Kiesecker, J. M., H. Copeland, A. Pocewicz, et al. 2009b. A frame-

Knick, S. T., S. E. Hanser, R. F. Miller, et al. 2011. Ecological influ-

work for implementing biodiversity offsets: Selecting sites and

ence and pathways of land use in sagebrush. In S. T. Knick and J.

determining scale. BioScience 59:7784.


Kilpatrick, A. M., C. M. Gillin, and P. Daszak. 2009. Wildlife-livestock conflict: The risk of pathogen transmission from bison to
cattle outside Yellowstone National Park. J. Appl. Ecology 46:
47685.
Kimball, S. L., B. D. Bennett, and F. B. Salisbury. 1973. The growth
and development of montane species at near-freezing temperatures. Ecology 54:16673.

W. Connelly, eds., Greater Sage-grouse: Ecology and Conservation of a Landscape Species and Its Habitats, pp. 20351. Studies
in Avian Biology 38. Univ. California Press, Berkeley.
Knight, D. H. 1991. Pine forests: A comparative overview of eco
system structure and function. In N. Nakagoshi and F. B. Galley, eds., Coniferous Forest Ecology from an International
Perspective, pp. 12135. SPB Academic Publishing, The Hague.
Knight, D. H., and W. A. Reiners. 2000. Natural patterns in South-

Kimble, D. S., D. B. Tyers, J. Robison-Cox, and B. F. Sowell. 2011.

ern Rocky Mountain landscapes and their relevance to forest

Aspen recovery since wolf reintroduction on the northern

management. In R. L. Knight, F. W. Smith, S. W. Buskirk, et al.,

Yellowstone winter range. Rangeland Ecol. Manage. 64:11930.

eds., Forest Fragmentation in the Southern Rocky Mountains,

King, D. A., D. M. Bachelet, and A. J. Symstad. 2013. Vegetation

pp. 132. Univ. Press Colorado, Niwot.

projections for Wind Cave National Park with three future cli-

Knight, D. H., B. S. Rogers, and C. R. Kyte. 1977. Understory plant

mate scenarios. Natural Resource Technical Report NPS/WICA/

growth in relation to snow duration in Wyoming subalpine for-

NRTR-2013/681. Wind Cave National Park, Hot Springs, South


Dakota.
Kipfmueller, K. F., and W. L. Baker. 2000. A fire history of a subalpine forest in south-eastern Wyoming. J. Biogeog. 27:7185.

est. Bull. Torrey Bot. Club 104:31419.


Knight, D. H., S. W. Running, and T. J. Fahey. 1985. Water and nutrient outflow from contrasting lodgepole pine forests in Wyoming. Ecol. Monogr. 55:2948.

Kirkham, D. R., and H. G. Fisser. 1972. Rangeland relations and

Knight, D. H., J. B. Yavitt, and G. D. Joyce. 1991. Water and nitrogen

harvester ants in north-central Wyoming. J. Range Manage. 25:

outflow from lodgepole pine forest after two levels of tree mor-

5560.
Kirol, C. P., J. L. Beck, J. B. Dinkins, and M. R. Conover. 2012.

tality. Forest Ecol. Manage. 46:21525.


Knight, S. H. 1974. Geologic history of Wyoming landscapes. In

Microhabitat selection for nesting and brood-rearing by the

M. Wilson, ed., Applied Geology and Archaeology: The Holo-

greater sage-grouse in xeric big sagebrush. Condor 114:7589.

cene History of Wyoming, pp. 17. Rep. Invest. 10. Geological

Kittel, G. 1994. Montane riparian vegetation in relation to elevation and geomorphology along the Cache la Poudre River, Colorado. M.S. thesis, Univ. Wyoming, Laramie.
Kleiner, E. F., and K. T. Harper. 1977. Soil properties in relation to
cryptogamic groundcover in Canyonlands National Park. J.
Range Manage. 30:2025.
Klutsch, J. G., J. F. Negron, S. L. Costello, et al. 2009. Stand characteristics and downed woody debris accumulations associated
with a mountain pine beetle (Dendroctonus ponderosae Hopkins)
outbreak in Colorado. Forest Ecol. Manage. 258:64149.
Klutsch, J. G., M. A. Battaglia, D. R. West, et al. 2011. Evaluating potential fire behavior in lodgepole pinedominated forests after
a mountain pine beetle epidemic in north-central Colorado.
West. J. Appl. Forest. 26:1019.
Knapp, A. K., and T. R. Seastedt. 1986. Detritus accumulation limits
productivity of tallgrass prairie. BioScience 36:66268.
Knapp, A. K., and W. K. Smith. 1981. Water relations and succession in subalpine conifers in southeastern Wyoming. Bot. Gaz.
142:50211.
. 1982. Factors influencing understory seedling establish-

Survey of Wyoming, Laramie.


. 1990. Illustrated geologic history of the Medicine Bow
Mountains and adjacent areas, Wyoming. Memoir 4. Geological Survey of Wyoming, Laramie.
Knopf, F. L., and M. L. Scott. 1990. Altered flows and created landscapes in the Platte River headwaters 18401990. In J. M. Sweeney, ed., Management of Dynamic Ecosystems, pp. 4770. North
Central Section, Wildlife Society, West Lafayette, Indiana.
Knowles, P., and M. C. Grant. 1983. Age and size structure analyses
of Engelmann spruce, ponderosa pine, lodgepole pine, and limber pine in Colorado. Ecology 64:19.
Knutson, R. M. 1981. Flowers that make heat while the sun shines.
Nat. Hist. (Oct.) 90:7581.
Kolb, K. J., and J. S. Sperry. 1999a. Differences in drought adaptation between subspecies of sagebrush (Artemisia tridentata).
Ecology 80:237384.
. 1999b. Transport constraints on water use by the Great
Basin shrub, Artemisia tridentata. Plant Cell Environ. 22:92535.
Kolm, K. E. 1975. Selenium in soils of the lower Wasatch formation,
Campbell County, Wyoming: Geochemistry, distribution, and

ment of Engelmann spruce (Picea engelmannii) and subalpine

environmental hazards. M.S. thesis, Univ. Wyoming, Laramie.

fir (Abies lasiocarpa) in southeastern Wyoming. Can. J. Bot. 60:

Komarek, E. V., Sr. 1964. The natural history of lightning. In Third

275361.
Knapp, P. A., and P. T. Soul. 2008. Use of atmospheric CO2-sensitive
trees may influence dendroclimatic reconstructions. Geophys.
Res. Lett. 35:L24703. doi: 10.1029/2008l035664.
Knick, S. T., and J. W. Connelly, eds. 2011. Greater Sage-grouse: Ecology and Conservation of a Landscape Species and Its Habitats.
Studies in Avian Biology 38. Univ. California Press, Berkeley.

Annual Tall Timber Fire Ecol. Conf., pp. 13983. Tallahassee,


Florida.
Korb, N. T. 2005. Historical fire regimes and structures of Douglasfir forests in the Centennial Valley of southwest Montana. M.S.
thesis, Colorado State Univ., Fort Collins.
Krner, C. 2012. Alpine Treelines: Functional Ecology of the Global
High Elevation Tree Limits. Springer, New York.

365

366 References
Kornfeld, M., and A. J. Osborn, eds. 2003. Islands on the Plains:

Lanier, J. E. 1978. Population dynamics of Rocky Mountain Doug-

Ecological, Social, and Ritual Use of Landscapes. Univ. Utah

las fir (Psuedotsuga menziesii var. glauca) along an elevational

Press, Salt Lake City.

gradient in the front range of Colorado. M.A. thesis, Univ. Colo-

Kornfeld, M., G. C. Frison, and M. L. Larson, et al. 2010. Prehistoric


Hunter-gatherers of the High Plains and Rockies. Left Coast
Press, Walnut Creek, California.
Koske, R. E., and W. R. Polson. 1984. Are VA mycorrhizae required
for sand dune stabilization? BioScience 34:42024.
Krause, T. R., and C. Whitlock. 2013. Climate and vegetation
change during the late-glacial/early Holocene transition inferred from multiple proxy records from Blacktail Pond, Yellowstone National Park, USA. Quaternary Res. 79:391402.
Krebill, R. G. 1972. Mortality of aspen on the Gros Ventre elk winter range. U.S. Forest Ser. Res. Paper INT-129.
Krueger, K. 1986. Feeding relationships among bison, pronghorn,
and prairie dogs: An experimental analysis. Ecology 67:76070.
Krzyszowska-Waitkus, A., G. F. Vance, and C. M. Preston. 2006. Influence of coarse wood and fine litter on forest organic matter
composition. Can. J. Soil Sci. 86:3546.
Kulakowski, D., and D. Jarvis. 2011. The influence of mountain
pine beetle outbreaks and drought on severe wildfires in northwestern Colorado and southern Wyoming: A look at the past
century. Forest Ecol. Manage. 262:168696.
Kulakowski, D., T. T. Veblen, and S. Drinkwater. 2004. The persistence of quaking aspen (Populus tremuloides) in the Grand Mesa
Area, Colorado. Ecol. Appl. 14:160314.

rado, Boulder.
Laramie Chamber of Commerce. 1913. Resources of Albany County.
Laramie, Wyoming.
Larson, E. R. 2011. Influences of the biophysical environment on
blister rust and mountain pine beetle, and their interactions, in
whitebark pine forests. J. Biogeogr. 38:45370.
Larson, F. 1940. The role of bison in maintaining the short grass
plains. Ecology 21:11321.
Larson, G. E., and J. R. Johnson. 1999. Plants of the Black Hills and
Bear Lodge Mountains. South Dakota Agric. Exp. Sta. B732,
Brookings.
Larson, T. A. 1977. Wyoming: A Bicentennial History. Norton, New
York.
. 1978. History of Wyoming, 2nd edition. Univ. Nebraska
Press, Lincoln.
Lauenroth, W. K., and I. C. Burke, eds. 2008. Ecology of the Shortgrass Steppe. Oxford Univ. Press, Oxford.
Lauenroth, W. K., J. L. Dodd, and P. L. Sims. 1978. The effects of
water- and nitrogen-induced stresses on plant community
structure in a semiarid grassland. Oecologia 36:21122.
. 1979. Response of native grassland legumes to water and
nitrogen treatments. J. Range Manage. 32:29294.

Kulakowski, D., M. W. Kaye, and D. M. Kashian. 2013. Long-term

Lauenroth, W. K., D. G. Milchunas, J. L. Dodd, et al. 1994. Effects

aspen cover change in the western U.S. Forest Ecol. Manage.

of grazing on ecosystems of the Great Plains. In M. Vavra, W.A.

299:5259.
Kulmatiski, A., K. H. Beard, L. A. Meyerson, et al. 2010. Nonnative
Phragmites australis invasion into Utah wetlands. West. N. Am.
Nat. 70:54152.
Kunzig, R. 2011. World without ice. Natl. Geogr. 220 (October):
90109.
Kupfer, J. A., and D. M. Cairns. 1996. The suitability of montane
ecotones as indicators of global climatic change. Prog. Phys.
Geog. 20:25372.
Kuvlesky, Jr., W. P., L. A. Brennan, M. L. Morrison, et al. 2007. Wind
energy development and wildlife conservation: Challenges and
opportunities. J. Wildlife Manage. 71:247898.
Kwon, H., E. Pendall, B. E. Ewers, et al. 2008. Spring drought regulates summer net ecosystem CO2 exchange in a sagebrushsteppe ecosystem. Agric. Forest. Meteorol. 148:38191.
Kyte, C. R. 1975. An autecological study of Vaccinium scoparium in
the Medicine Bow Mountains, Wyoming. M.S. thesis, Univ.
Wyoming, Laramie.
Ladenburger, C. G., A. L. Hild, D. J. Kazmer, and L. C. Munn. 2006.
Soil salinity patterns in Tamarix invasions in the Bighorn Basin,
Wyoming, USA. J. Arid Environ. 65:11128.
Lageson, D. R., and D. R. Spearing. 1988. Roadside Geology of Wyoming. Mountain Press Publishers, Missoula, Montana.
Lang, R. L. 1973. Vegetation changes between 1943 and 1965 on
the shortgrass plains of Wyoming. J. Range Manage. 26:4079.
Langenheim, J. H. 1962. Vegetation and environmental patterns

Laycock, and R. D. Pieper, eds., Ecological Implications of Livestock Herbivory in the Great Plains of the United States, pp.69
100. Soc. Range Manage., Denver.
Lauenroth, W. K., D. G. Milchunas, O. E. Sala, et al. 2008. Net primary production in the shortgrass steppe. In W. K. Lauenroth
and I. C. Burke, eds., Ecology of the Shortgrass Steppe, pp. 270
305. Oxford Univ. Press, Oxford.
Lauenroth, W. K., R. L. Dougherty, and J. S. Singh. 2009. Precipitation event size controls on long-term abundance of Opuntia
polyacantha (plains pricklypear) in Great Plains grasslands.
Great Plains Res. 19:5564.
Launchbaugh, K. L., F. D. Provenza, and J. A. Pfister. 2001. Herbivore response to anti-quality factors in forages. J. Range Manage. 54:43140.
Lawrence, D. B., and E. G. Lawrence. 1984. Historic lower landslide
of the Gros Ventre Valley. Naturalist 35 (winter):2428.
Laycock, W. A. 1958. The initial pattern of revegetation of pocket
gopher mounds. Ecology 39:34651.
. 1991. Stable states and thresholds of range condition on
North American rangelands: A viewpoint. J. Range Manage. 44:
42733.
League, K., and T. Veblen. 2006. Climatic variability and episodic
Pinus ponderosa establishment along the forest-grassland ecotones of Colorado. Forest Ecol. Manage. 228 98107.
Leatherman, D. A., J. W. Brewer, and R. E. Stevens. 2009. Western

in the Crested Butte area, Gunnison County, Colorado. Ecol.

Spruce Budworms. Colorado State Univ. Coop. Ext., Insect Se-

Monogr. 32:24985.

ries 5.543. Fort Collins, Colorado.

References
LeBeau, C. W., J. L. Beck, G. D. Johnson, and M. J. Holloran. 2014.

Lesica, P., S. V. Cooper, and G. Kudray. 2007. Recovery of big sage-

Short-term impacts of wind energy development on greater

brush following fire in southwest Montana. Rangeland Ecol.

sage-grouse fitness. J. Wildlife Manage. 78 [in press].


Lee, J. 1916. Diary of Rev. Jason Lee. Quart. Oregon Hist. Soc. 17:
116146.
Leffler, A. J., C. Y. Ivans, R. J. Ryel, and M. M. Caldwell. 2004. Gas
exchange and growth responses of the desert shrubs Artemisia
tridentata and Chrysothamnus nauseosus to shallow- vs. deep-soil
water in a glasshouse experiment. Environ. Exp. Bot. 51:919.

Manage. 60:26169.
Lesser, M. R., and S. T. Jackson. 2012. Making a stand: Five centuries of population growth in colonizing populations of Pinus
ponderosae. Ecology 93:107181.
Lewis, D. 2011. A Field Guide to Reptiles and Amphibians of Wyoming. Wyoming Naturalist, Douglas, Wyoming.
Licciardi, J. M., and K. L. Pierce. 2008. Cosmogenic exposure-age

Leffler, A. J., M. S. Peek, R. J. Ryel, et al. 2005. Hydraulic redistri-

chronologies of Pinedale and Bull Lake glaciations in greater

bution through the root systems of senesced plants. Ecology

Yellowstone and the Teton Range, USA. Quaternary Sci. Rev.

86:63342.
Leger, E. A. 2008. The adaptive value of remnant native plants in
invaded communities: An example from the Great Basin. Ecol.
Appl. 18:122635.
Lemly, J. M. 2007. Fens of Yellowstone National Park, USA: Regional
and local controls over plant species distribution. Ph.D. diss.,
Colorado State Univ., Fort Collins.
Lemly, J. M., and D. J. Cooper. 2011. Multiscale factors control community and species distribution in mountain peatlands. Botany
89:689713.
Lentile, L. B., F. W. Smith, and W. D. Shepperd. 2006. Influence of
topography and forest structure on patterns of mixed severity
fire in ponderosa pine forests of the South Dakota Black Hills,
USA. Int. J. Wildlife Fire 15:55766.

27:81431.
Ligon, J. D. 1978. Reproductive interdependence of pinion jays and
pinion pines. Ecol. Monogr. 48:11126.
Lillegraven, J. A., and L. M. Ostresh, Jr. 1988. Evolution of Wyomings early Cenozoic topography and drainage patterns. Natl.
Geogr. Res. 4:30327.
Linhart, Y. B., J. B. Mitton, K. C. Sturgeon, and M. L. Davis. 1981.
Genetic variation in space and time in a population of ponderosa pine. Heredity 46:40226.
Liptzin, D., and T. R. Seastedt. 2010. Regional and local patterns
of soil nutrients at Rocky Mountains treelines. Geoderma 160:
20817.
Lisenbee, A., F. Karner, E. Kashbaugh, et al. 1981. Geology of the
Tertiary intrusive province of the Northern Black Hills, South

Leopold, A. 1938. Engineering and conservation. A lecture deliv-

Dakota and Wyoming. In F. J. Rich, ed., Geology of the Black

ered to the University of Wisconsin College of Engineering in

Hills, South Dakota and Wyoming, pp. 33105. Am. Geol. Inst.,

1938. Reprinted in S. L. Flader and J. B. Caldicott, eds. 1991. The


River of the Mother of God and Other Essays by Aldo Leopold,
pp. 24954. Univ. Wisconsin Press, Madison.
Leopold, A. S., S. A. Cain, C. M. Cottam, et al. 1963. Wildlife management in the national parks. Trans. N. Am. Wildlife Nat. Res.
Conf. 28:2845.
Leopold, E. B., and M. F. Denton. 1987. Comparative age of grassland and steppe east and west of the Northern Rocky Mountains. Ann. Missouri Bot. Garden 74:84167.
Lepper, M. G. 1974. Pinus flexilis and its environmental relationships. Ph.D. diss., Univ. California, Davis.

Falls Church, Virginia.


Litaor, M. I., T. R. Seastedt, and D. A. Walker. 2002. Spatial analysis of selected soil attributes across an alpine topographic/snow
gradient. Landscape Ecol. 17:7185.
Litaor, M. I., M. Williams, and T. R. Seastedt. 2008. Topographic controls on snow distribution, soil moisture, and species diversity
of herbaceous alpine vegetation, Niwot Ridge, Colorado. J. Geophys. Res. Biogeosci. 113:G02008. doi: 10.1029/2007JG000419.
Littell, J. S. 2002. Determinants of fire regime variability in lower
elevation forests of the northern Greater Yellowstone Ecosystem. M.S. thesis, Montana State Univ., Bozeman.

Lepper, M. G., and M. Fleschner. 1977. Nitrogen fixation by Cerco

Litton, C. M., M. G. Ryan, D. H. Knight, and P. D. Stahl. 2003. Soil-

carpus ledifolius (Rosaceae) in pioneer habitats. Oecologia 27:

surface carbon dioxide efflux and microbial biomass in relation

33338.

to tree density 13 years after a stand replacing fire in a lodgepole

Leppi, J. C., T. H. DeLuca, S. W. Harrar, and S. W. Running. 2011.

pine ecosystem. Global Change Biol. 9:68096.

Impacts of climate change on August stream discharge in

Lockwood, J. A. 2004. Locust: The Devastating Rise and Mysterious

the central Rocky Mountains. Climate Change. doi: 10.1007/

Disappearance of the Insect that Shaped the American Frontier.

s10584-011-0235-1.

Basic Books, New York.

Lesica, P., and S. Miles. 2001a. Natural history and invasion of

Lockwood, J. A., T. J. McNary, J. C. Larsen, and J. Cole. 1993. Distri-

Russian olive along eastern Montana rivers. West. N. Am. Nat.

bution atlas for grasshoppers and the Mormon cricket in Wyo-

61:110.
. 2001b. Tamarisk growth at the northern margin of its naturalized range in Montana, USA. Wetlands 21:24046.
. 2004a. Beavers indirectly enhance the growth of Russian
olive and tamarisk along eastern Montana rivers. West. N. Am.
Nat. 64:93100.

ming, 19881992. Wyoming Agric. Exp. Sta. Bull. 976, Laramie.


Lockwood, J. A., S. P. Schell, R. N. Foster, et al. 2000. Reduced agentarea treatments (RAATs) for management of rangeland grasshoppers: Efficacy and economics under operational conditions.
Int. J. Pest Manage. 46:2942.
Lockwood, J. A., R. Anderson-Sprecher, and S. P. Schell. 2002.

. 2004b. Ecological strategies for managing tamarisk on the

When less is more: Optimization of reduced agent-area treat-

C. M. Russell National Wildlife Refuge, Montana, USA. Biol.

ments (RAAT) for management of rangeland grasshoppers.

Conserv. 119:53543.

Crop Protect. 21:55162.

367

368 References
Loffler, J., K. Anschlag, B. Baker, et al. 2011. Mountain ecosystem
response to global change. Erdkunde 65:189213.
Logan, J. A., and J. A. Powell. 2001. Ghost forests, global warming,
and the mountain pine beetle (Coleoptera: Scolytidae). Am.
Entomol. 47:16072.

Lundberg, C. E. 1977. The local composition and distribution of


vegetation in the soda ash area of Sweetwater County, Wyoming. M.S. thesis, Univ. Wyoming, Laramie.
Lundquist, J. D., M. D. Dettinger, I. T. Steward, and D. R. Cayan.
2009. Variability and trends in spring runoff in the western

Logan, J. A., W. W. Macfarlane, and L. Willcox. 2010. Whitebark

United States. In F. H. Wagner, ed., Climate Warming in West-

pine vulnerability to climate-driven mountain pine beetle

ern North America: Evidence and Environmental Effects, pp.

disturbance in the Greater Yellowstone Ecosystem. Ecol. Appl.


20:895902.
Lohman, K. 2004. Wildlife use of riverine wetlands. In M. C. Mc
Kinstry, S. H. Anderson, and W. A. Hubert, eds., Wetland and
Riparian Areas of the Intermountain West: Their Ecology and
Management, pp. 7486. Univ. Texas Press, Austin.
Lommasson, T. 1948. Succession in sagebrush. J. Range Manage.
1:1921.
Longland, W. S., and S. L. Bateman. 2002. The ecological value of
shrub islands on disturbed sagebrush rangelands. J. Range Manage. 55:57175.
Loope, L. L., and G. E. Gruell. 1973. The ecological role of fire in
Jackson Hole, northwestern Wyoming. Quaternary Res. 3:
42543.
Lotan, J. E. 1964. Regeneration of lodgepole pine: A study of slash
disposal and cone opening. U.S. Forest Ser. Res. Note INT-16.
. 1975. The role of cone serotiny in lodgepole pine forests. In
D. M. Baumgartner, ed., Management of Lodgepole Pine Ecosystems, pp. 47195. Washington State Univ. Coop. Ext. Serv.,
Pullman.
Lotan, J. E., and D. A. Perry. 1983. Ecology and Regeneration of
Lodgepole Pine. USDA Handbook 606. U.S. Dept. Agric., Washington, D.C.
Lovaas, A. L. 1976. Introduction of prescribed burning to Wind
Cave National Park. Wildlife Soc. Bull. 4:6973.
Love, J. D., and A. C. Christensen. 1985. Geologic Map of Wyoming. U.S. Geological Survey, Denver.
Love, J. D., J. C. Reed, Jr., and K. L. Pierce. 2003. Creation of
the Teton Landscape. Grand Teton Nat. Hist. Assoc., Moose,
Wyoming.
Lovvorn, J. R., and E. A. Hart. 2004. Irrigation, salinity, and landscape patterns of natural palustrine wetlands. In M. C. Mc

6376. Univ. Utah Press, Salt Lake City.


Lyford, M. E., J. L. Betancourt, and S. T. Jackson. 2002. Holocene
vegetation and climate history of the northern Bighorn Basin,
southern Montana. Quaternary Res. 58:17181.
Lyford, M. E., S. T. Jackson, J. L. Betancourt, and S. T. Gray. 2003.
Influence of landscape structure and climate variability on a
late Holocene plant migration. Ecol. Monogr. 73:56783.
Lynch, E. A. 1998. Origin of a park-forest vegetation mosaic in the
Wind River Range, Wyoming. Ecology 79:132038.
Lynch, H. J., R. A. Renkin, R. L. Crabtree, and P. R. Moorcroft. 2006.
The influence of previous mountain pine beetle (Dendroctonus
ponderosae) activity on the 1988 Yellowstone fires. Ecosystems
9:131827.
MacCracken, J. G., D. W. Uresk, and R. M. Hansen. 1983. Plant community variability on a small area in southeastern Montana.
Great Basin Nat. 43:66068.
MacDonald, L. H., and J. D. Stednick. 2003. Forests and water: A
state-of-the-art review for Colorado. Completion Report 196.
Colorado Water Resour. Res. Inst., Fort Collins.
Mack, R. N. 2011a. Cheatgrass. In D. Simberloff and M. Rejmanek,
eds., Encyclopedia of Biological Invasions, pp. 10813. Univ.
California Press, Berkeley.
. 2011b. Fifty years of waging war on cheatgrass: Research
advances while meaningful control languishes. In D. M. Richardson, ed., Fifty Years of Invasion Ecology, pp. 25365. WileyBlackwell, West Sussex, UK.
Mack, R. N., and J. N. Thompson. 1982. Evolution in steppe with
few large, hoofed mammals. Am. Nat. 119:75773.
MacMahon, J. A., and D. C. Anderson. 1982. Subalpine forests: A
world perspective with emphasis on western North America.
Prog. Phys. Geog. 6:368425.
Madson, C. 2006a. Gas fields and wildlife. Wyoming Wildlife 70:
1524.

Kinstry, W. A. Hubert, and S. H. Anderson, eds., Wetland and

. 2006b. In search of compromise. Wyoming Wildlife 70:4.

Riparian Areas of the Intermountain West: Ecology and Man-

. 2007. Making a change. Wyoming Wildlife 71:4.

agement, pp. 10529. Univ. Texas Press, Austin.

. 2012. Going with the flow. Wyoming Wildlife 76:3135.

Lovvorn, J. R., W. M. Wollheim, and E. A. Hart. 1999. High plains

Major, J., and M. Rejmanek. 1992. Amelanchier alnifolia vegetation

wetlands of southeast Wyoming. Salinity, vegetation and in-

in eastern Idaho, U.S.A. and its environmental relationships.

vertebrate communities. In D. P. Batzer, R. B. Rader, and S. A.

Vegetatio 98:14156.

Wissinger, eds., Invertebrates in Freshwater Wetlands of North

Malanson, G. P., and D. R. Butler. 1984. Transverse pattern of veg-

America: Ecology and Management, pp. 60334. John Wiley

etation on avalanche paths in the northern Rocky Mountains,

and Sons, Hoboken, New Jersey.


Ludlow, W. 1875. Report of a reconnaissance of the Black Hills of

Montana. Great Basin Nat. 44:45358.


Malanson, G. P., D. R. Butler, D. B. Fagre, et al. 2007. Alpine treeline of

Dakota made in the summer of 1874. U.S. Army, Dept. of Engi-

western North America: Linking organism-to-landscape dynam-

neers. U.S. Gov. Printing Office, Washington, D.C.

ics. Phys. Geogr. 28:37896. doi: 0.2747/0272-3646.28.5.378.

Ludwig, J. A. 1969. Environmental interpretation of foothill grass-

Manier, D. J., and N. T. Hobbs. 2007. Large herbivores in sagebrush

land communities of northern Utah. Ph.D. diss., Univ. Utah,

steppe ecosystems: Livestock and wild ungulates influence

Salt Lake City.

structure and function. Oecologia 152:73950.

References
Manier, D. J., and R. D. Laven. 2002. Changes in landscape patterns

Mattson, D. J., S. R. Podruzny, and M. A. Haroldson. 2002. Con-

associated with the persistence of aspen (Populus tremuloides

sumption of fungal sporocarps by Yellowstone grizzly bears.

Michx.) on the western slope of the Rocky Mountains, Colorado. Forest Ecol. Manage. 167:26384.
Mann, C. 2005. 1491: New Revelations of the Americas before Columbus. Knopf, New York.

Ursus 13:95103.
Mayer, J. H. 2003. Paleoindian geoarchaeology and paleoenvironments of the western Killpecker Dunes, Wyoming, USA. Geo
archaeol. Int. J. 18:3569.

Marcott, S. A., J. D. Shakun, P. U. Clark, and A. C. Mix. 2013. A

Mayer, J. H., and S. A. Mahan. 2004. Late Quaternary stratigraphy

reconstruction of regional and global temperature for the past

and geochronology of the western Killpecker Dunes, Wyoming,

11,300 years. Science 339:11982001.


Marcus, W. A., ed. 2012. Atlas of Yellowstone. Univ. California
Press, Berkeley.
Marr, J. W. 1977. The development and movement of tree islands
near the upper limit of tree growth in the southern Rocky
Mountains. Ecology 58:115964.
Marriott, H. J. 2012. Survey and mapping of Black Hills montane
grasslands. Report. South Dakota Dept. Game, Fish and Parks,
Pierre.
Marriott, H. J., and D. Faber-Langendoen. 2000. Black Hills community report: Plant community descriptions (three reports
on upland and wetland vegetation). Nature Conservancy and
Association for Biodiversity Information. http://conserveonline
.org/library/.
Marriott, H. J., D. Faber-Langendoen, A. McAdams, et al. 1999.
Black Hills community inventory. Final report. Nature Conservancy, Midwestern Resource Office, Minneapolis, Minnesota.
Marshall, K. N., N. T. Hobbs, and D. J. Cooper. 2013. Stream hydrology limits recovery of riparian ecosystems after wolf reintroduction. Proc. Royal Soc. B 280:20122977.
Marston, R. A., and J. E. Anderson. 1991. Watersheds and vegetation
of the Greater Yellowstone Ecosystem. Conserv. Biol. 5:33846.
Marston, R. A., and D. J. Brosz. 1990. Drainage basins. In L. M.
Ostresh, Jr., R. A. Marston, and W. M. Hudson, eds., Wyoming
Water Atlas, pp. 4647. Wyoming Water Development Commission, Cheyenne.
Martin, C. W. 1992. Late Holocene alluvial chronology and climate
change in the central Great-Plains. Quaternary Res. 37:31522.
Martner, B. E. 1986. Wyoming Climate Atlas. Univ. Nebraska Press,
Lincoln.
Maser, C., J. M. Trappe, and R. A. Nussbaum. 1978. Fungussmall
mammal interrelationships with emphasis on Oregon coniferous forests. Ecology 59:799809.

USA. Quaternary Res. 61:7284.


McBride, J. R., and J. Strahan. 1984. Establishment and survival
of woody riparian species on gravel bars of an intermittent
stream. Am. Midl. Nat. 112:23545.
McCulley, R. L., E. G. Jobbagy, W. T. Pockman, and R. B. Jackson.
2004. Nutrient uptake as a contributing factor for deep rooting
in arid and semi-arid ecosystems. Oecologia 141:62028.
McDonald, G. I., B. A. Richardson, P. J. Zambino, et al. 2006.
Pedicularis and Castilleja are natural hosts of Cronartium ribicola
in North America: A first report. Forest Pathol. 36:7382.
McDonough, W. T. 1985. Sexual reproduction, seeds, and seedlings. In N. V. DeByle and R. P. Winokur, eds., Aspen: Ecology
and Management in the Western United States. U.S. Forest Ser.
Gen. Tech. Rep. RM-119.
McDonough, W. T., and R. O. Harniss. 1974. Seed dormancy in Artemisia tridentata Nutt. subspecies vaseyana Rydb. Northwest. Sci.
48:1720.
McFaul, M. 1979. A geomorphic and pedological interpretation of
the mima-mound prairies, south Puget lowland, Washington
state. M.A. thesis, Univ. Wyoming, Laramie.
McGinnies, W. J. 1960. Effect of mima-type microrelief on herbage
production of five seeded grasses in western Colorado. J. Range
Manage. 13:23134.
McGinnies, W. J., L. W. Osburn, and W. A. Berg. 1976. Plant-soilmicrosite relationships on a saltgrass meadow. J. Range Manage.
29:395400.
McGinnies, W. J., H. L. Shantz, and W. G. McGinnies. 1991.
Changes in vegetation and land use in eastern Colorado: A
photog raphic study. U.S. Agric. Res. Ser. ARS-85.
McGranahan, D. A., D. M. Engle, S. D. Fuhlendorf, et al. 2012. Spatial heterogeneity across five rangelands managed with pyricherbivory. J. Appl. Ecol. 49:90310.

Maser, C., R. F. Tarrant, J. M. Trappe, and J. F. Franklin, eds. 1988.

McInerney, F.A., and S.L. Wing. 2011. The Paleocene-Eocene ther-

From the forest to the sea: A story of fallen trees. U.S. Forest Ser.

mal maximum: A perturbation of carbon cycle, climate, and

Gen. Tech. Rep. PNW-GTR-229.

biosphere with implications for the future. Annu. Rev. Earth

Mason, J. A., J. B. Swinehart, R. R. Hanson, et al. 2011. Late Pleistocene dune activity in the central Great Plains, USA. Quaternary
Sci. Rev. 30:385870.
Mason, M. K. 1987. LaramieGem City of the Plains. Curtis Media,
Dallas.

Planetary Sci. 39:489516.


McIntire, P. W. 1984. Fungus consumption by the Siskiyou chipmunk within a variously treated forest. Ecology 65:13746.
McIntosh, A. C. 1931. A botanical survey of the Black Hills of South
Dakota. Black Hills Engr. 19:159276.

Matson, P. A., C. Volkmann, K. Coppinger, and W. A. Reiners. 1991.

McKinstry, M. C., and S. H. Anderson. 1999. Attitudes of private-

Annual nitrous oxide flux and soil nitrogen characteristics in

and public-land managers in Wyoming, USA, toward beaver.

sagebrush steppe ecosystems. Biogeochem. 14:12.


Mattson, D. J. 1984. Classification and environmental relationships

Environ. Manage. 23:95101.


McKinstry, M. C., S. H. Anderson, and W. A. Hubert, eds. 2004.

of wetland vegetation in central Yellowstone National Park.

Wetland and Riparian Areas of the Intermountain West: Their

M.S. thesis, Univ. Idaho, Moscow.

Ecology and Management. Univ. Texas Press, Austin.

369

370 References
McNaughton, S. J. 1976. Serengeti migratory wildebeest: Facilitation of energy flow by grazing. Science 191:9294.
. 1985. Ecology of a grazing ecosystem: The Serengeti. Ecol.
Monogr. 55:25994.
Meagher, M., and D. B. Houston. 1998. Yellowstone and the biology
of time: Photographs across a century. Univ. Oklahoma Press,
Norman.
Mealor, B. A., and A. L. Hild. 2007. Post-invasion evolution of native plant populations: A test of biological resilience. Oikos
116:14931500.
Mears, A. I. 1975. Dynamics of dense-snow avalanches interpreted
from broken trees. Geology 3:52123.

study of regulated and unregulated streams in the Green River


Basin, USA. Regulated Rivers Res. Manage. 16:54364.
Merritt, D. M., and L. N. Poff. 2010. Shifting dominance of riparian
Populus and Tamarix along gradient of flow alteration in western North American rivers. Ecol. Appl. 20:13552.
Meyer, C. B., D. H. Knight, and G. K. Dillon. 2005. Historic range
of variability for upland vegetation in the Bighorn National Forest, Wyoming. U.S. Forest Ser. RMRS-GTR-140.
Meyer, S. E., and S. B. Monsen. 1991. Habitat-correlated variation in
mountain big sagebrush (Artemisia tridentata ssp. vaseyana) seed
germination patterns. Ecology 72:73942.
Meyer, S. E., S. C. Garvin, and J. Beckstead. 2001. Factors mediat-

Mears, B., Jr. 1981. Periglacial wedges and the late Pleistocene

ing cheatgrass invasion of intact salt desert shrubland. In E. D.

environment of Wyomings intermontane basins. Quaternary

McArther and D. J. Fairbanks, compilers, Shrubland ecosystem

Res. 15:17198.

genetics and biodiversity. Proceedings of a symposium, pp. 224

. 1987. Late Pleistocene periglacial wedge sites in Wyoming:


An illustrated compendium. Memoir 3. Geological Survey of
Wyoming, Laramie.
. 1991. Laramie Basin. In R. B. Morrison, ed., Quaternary
Nonglacial Geology: Conterminous U.S. The Geology of North
America, volume K-2, pp. 41827. Geol. Soc. Am., Boulder.

31. USDA RMRS-P-21. U.S. Dept. Agric., Washington, D.C.


Meyer, S. E., D. Quinney, D. Nelson, and J. Weaver. 2007. Impact
of the pathogen Pyrenophora semeniperda on Bromus tectorum
seedbank dynamics in North American cold deserts. Weed Res.
47:5462.
Middleton, A. D., M. J. Kaufman, D. E. McWhirter, et al. 2013a.

. 1993. Geomorphic history of Wyoming and high-elevation

Linking anti-predator behavior to prey demography reveals

erosion surfaces. In A. W. Snoke, J. R. Steidtmann, and S. M.

limited risk effects of an actively hunting large carnivore. Ecol.

Roberts, eds., Geology of Wyoming, pp. 60826. Memoir 5.


Geological Survey of Wyoming, Laramie.
. 1997. Glaciation of the Wyoming mountains and Pleistocene permafrost in the basin floors. Public Information Circular 38. Geological Survey of Wyoming, Laramie.
. 2001. Glacial records in the Medicine Bow Mountains and
Sierra Madre of southern Wyoming and adjacent Colorado,
with a travelers guide to their sites. Public Information Circular
41. Geological Survey of Wyoming, Laramie.
Mears, B., Jr., W. P. Eckerle, D. R. Gilmer, et al. 1986. A Geologic
Tour of Wyoming from Laramie to Lander, Jackson and Rock
Springs. Public Info. Circ. 27. Wyo. Geol. Survey, Laramie.
Medin, D. E. 1960. Physical site factors influencing annual production of true mountain mahogany, Cercocarpus montanus. Ecology 41:45460.

Lett. 16:102330.
. 2013b. Animal migration amid shifting patterns of phenology and predation: Lessons from a Yellowstone elk herd. Ecology
94:124556.
. 2013c. Rejoinder: Challenge and opportunity in the study
of ungulate migration amid environmental change. Ecology
94:128086.
Middleton, A. D., T. A. Morrison, J. K. Fortin, et al. 2013d. Grizzly
bear predation links the loss of native trout to the demography of migratory elk in Yellowstone. Proc. Royal Soc. B 280:
20130870. http://dx.doi.org/10.1098/rspb.2013.0870.
Mielke, H. W. 1977. Mound building by pocket gophers (Geo
myidae): Their impact on soils and vegetation in North America. J. Biogeogr. 4:17180.
Mihlbachler, B. S. 1986. Effects of prescribed burning on peren-

Mehringer, P. J., Jr., and P. E. Wigand. 1990. Comparison of late Holo

nial grass vigor and production, plant water stress, and soil ero-

cene environments from woodrat middens and pollen: Diamond

sion on cactus rangelands in east-central Wyoming. M.S. thesis,

Craters, Oregon. In J. L. Betancourt, T. R. VanDevender, and P. S.


Martin, eds., Packrat Middens: The Last Forty Thousand Years of
Biotic Change, pp. 294325. Univ. Arizona Press, Tucson.
Meinke, C. W., S. T. Knick, and D. A. Pyke. 2009. A spatial model
to prioritize sagebrush landscapes in the Intermountain West
(USA) for restoration. Restoration Ecol. 17:65259.
Meinzer, F. C., J. R. Brooks, S. Bucci, et al. 2004. Converging patterns of uptake and hydraulic redistribution of soil water in
contrasting woody vegetation types. Tree Physiol. 24:91928.
Mensing, S., S. Livingston, and P. Barker. 2006. Long-term fire his-

Univ. Wyoming, Laramie.


Mikkelson, K. M., E. R. Dickenson, R. M. Maxwell, et al. 2013a.
Water quality impacts from climate-induced forest die-off. Nat.
Climate Change 3:21822.
Mikkelson, K. M., R. M. Maxwell, R. M. Ferguson, et al. 2013b.
Mountain pine beetle infestation impacts: Modeling water and
energy budgets at the hill-slope scale. Ecohydrology 6:6472.
Milchunas, D. G., O. E. Sala, and W. K. Lauenroth. 1988. A generalized model of the effect of grazing by large herbivores on grassland community structure. Am. Nat. 132:87106.

tory in Great Basin sagebrush reconstructed from macroscopic

Milchunas, D. G., W. K. Lauenroth, P. L. Chapman, and M. K. Kazem-

charcoal in spring sediments, Newark Valley, Nevada. West. N.

pour. 1989. Effects of grazing, topography, and precipitation on

Am. Naturalist 66:6477.

the structure of a semiarid grassland. Vegetatio 80:1123.

Merritt, D. M., and D. J. Cooper. 2000. Riparian vegetation and

Milchunas, D. G., W. K. Lauenroth, I. C. Burke, and J. K. Detling.

channel change in response to river regulation: A comparative

2008. Effects of grazing on vegetation. In W. K. Lauenroth and

References
I. C. Burke, eds., Ecology of the Shortgrass Steppe, pp. 389446.
Oxford Univ. Press, Oxford.
Miles, S. R., and P. C. Singleton. 1975. Vegetational history of Cinnabar Park in Medicine Bow National Forest, Wyoming. Soil Sci.
Soc. Am. Proc. 39:12048.
Millar, C. I., N. L. Stephenson, and S. L. Stephens. 2007. Climate
change and forests of the future: Managing in the face of un
certainty. Ecol. Applic. 17:214551.
Miller, B. J., R. P. Reading, D. E. Biggins, et al. 2007. Prairie dogs: An
ecological review and current biopolitics. J. Wildlife Manage.
71:280110.
Miller, J. R., T. T. Schulz, N. T. Hobbs, et al. 1995. Changes in the

Mitton, J. B., and S. M. Ferrenberg. 2012. Mountain pine beetle


develops an unprecedented summer generation in response to
climate warming. Am. Nat. 179:E163E171.
Mock, K. E., B. A. Richardson, and P. G. Wolf. 2013. Molecular tools
and aspen management: A primer and prospectus. Forest Ecol.
Manage. 299:613.
Mogk, D., D. Henry, P. Mueller, and D. Foster. 2012. Origins of a
continent. Yellowstone Sci. 20:2232.
Moir, W. H. 1969. The lodgepole pine zone in Colorado. Am. Midl.
Nat. 81:8798.
Moir, W. H., S. G. Rochelle, and A. W. Schoettle. 1999. Microscale
patterns of tree establishment near upper treeline, Snowy

landscape structure of a southeastern Wyoming riparian zone

Range, Wyoming, U.S.A. Arct. Antarct. Alp. Res. 31:37988.

following shifts in stream dynamics. Biol. Conserv. 72:37179.

Molenda, O., A. Reid, and C. J. Lortie. 2012. The alpine cushion

Miller, R. C. 1979. Relationships between the vegetation and

plant Silene acaulis as a foundation species: A bugs-eye view

environmental characteristics associated with an alluvial val-

to facilitation and microclimate. PLoS ONE 7(5):e37223. doi:

ley floor. M.S. thesis, Univ. Wyoming, Laramie.


Miller, R. F., and L. I. Shultz. 1987. Development and longevity

10.1371/journal.pone.0037223.
Monson, R. K., R. Mullen, and W. D. Bowman. 2001. Plant nutrient

of ephemeral and perennial leaves on Artemisia tridentata ssp.

relations. In W. D. Bowman and T. R. Seastedt, eds., Structure

wyomingensis. Great Basin Nat. 47:22730.

and Function of an Alpine Ecosystem, pp. 198221. Oxford

Miller, R. F., T. Svejcar, and N. West. 1993. Implications of livestock


grazing in the intermountain sagebrush region: Plant composi-

Univ. Press, New York.


Mooney, H. A., and W. D. Billings. 1960. The annual carbohydrate

tion. In M. Vavra, W. A. Laycock, and R. D. Pieper, eds., Ecologi-

cycle of alpine plants as related to growth. Am. J. Bot. 47:59498.

cal Implications of Livestock Herbivory in the West. Soc. Range

Moore, C. T. 1972. Man and fire in the central North American

Manage., Denver.
Miller, S. L., and E. B. Allen. 1992. Mycorrhizae, nutrient trans-

grassland 15351890: A documentary historical geography.


Ph.D. diss., Univ. California, Los Angeles.

location and interaction between plants. In M. F. Allen, ed.,

Moore, J. C., J. Sipes, A. A. Whittemore-Olson, et al. 2008. Trophic

Mycorrhizal Functioning: An Integrative Plant-Fungal Process,

structure and nutrient dynamics of the belowground food web

pp.30132. Chapman and Hall, New York.

within the rhizosphere of the shortgrass steppe. In W. K. Lauen-

Miller, W. B. 1964. An ecological study of the mountain-mahogany


community. M.S. thesis, Univ. Wyoming, Laramie.

roth and I. C. Burke, eds., Ecology of the Shortgrass Steppe,


pp.24869. Oxford Univ. Press, Oxford.

Millspaugh, S. H., C. Whitlock, and P. J. Bartlein. 2000. Variations

Morehouse, K., T. Johns, J. Kaye, and M. Kaye. 2008. Carbon and

in fire frequency and climate over the past 17,000 yr in central

nitrogen cycling immediately following bark beetle outbreaks

Yellowstone National Park. Geology 28:21114.

in southwestern ponderosa pine forests. Forest Ecol. Manage.

. 2004. Postglacial fire, vegetation, and climate history of the

255:26982708.

Yellowstone-Lamar and Central Plateau Provinces, Yellowstone

Morgan, J. A., D. G. Milchunas, D. R. LeCain, et al. 2007. Carbon di-

National Park. In L. L. Wallace, ed., After the Fires: The Ecology

oxide enrichment alters plant community structure and accel-

of Change in Yellowstone National Park, pp. 1028. Yale Univ.

erates shrub growth in the shortgrass steppe. Proc. Natl. Acad.

Press, New Haven.

Sci. USA 104:1472429.

Minckley, T. A., R. K. Shriver, and B. Shuman. 2012. Resilience and

Morgan, J. A., D. R. LeCain, E. Pendall, et al. 2011. C4 grasses pros-

regime change in a southern Rocky Mountain ecosystem dur-

per as carbon dioxide eliminates desiccation in warmed semi-

ing the past 17,000 years. Ecol. Monogr. 82:4968.


Mineau, M. M., C. V. Baxter, and A. M. Marcarelli. 2011. A nonnative riparian tree (Elaeagnus angustifolia) changed nutrient
dynamics in streams. Ecosystems 14:35365.
Minshall, G. W., J. T. Brock, and J. D. Varley. 1989. Wildfires and
Yellowstones stream ecosystems. BioScience 39:70715.
Minshall, G. W., C. T. Robinson, and D. E. Lawrence. 1997. Postfire responses of lotic ecosystems in Yellowstone National Park,
USA. Can. J. Fisheries Aquat. Sci. 54:250925.
Mitchell, J. E, and R. H. Hart. 1987. Winter of 198687: The death
knell of open range. Rangelands 9:38.
Mitsch, W. J., and J. G. Gosselink. 2007. Wetlands, 4th edition.
John Wiley and Sons, Hoboken, New Jersey.
Mitton, J. B. 1985. So grows the tree. Natural History (Jan.):5864.

arid grassland. Nature 476:2025.


Morris, L. E. 2013. The Perilous West. Rowman and Littlefield, Lanham, Maryland.
Morris, L. R., R. J. Ryel, and N. E. West. 2010. Can soil phytolith
analysis and charcoal be used as indicators of historic fire in the
pinyon-juniper and sagebrush steppe ecosystem types of the
Great Basin Desert, USA? Holocene 20:10514.
Morris, L. R., T. A Monaco, and R. L. Sheley. 2011. Land-use legacies
and vegetation recovery 90 years after cultivation in Great Basin
sagebrush ecosystems. Rangeland Ecol. Manage. 64:48897.
Mote, P. W. 2009. Variability and trends in mountain snowpacks
in western North America. In F. H. Wagner, ed., Climate Warming in Western North America: Evidence and Environmental
Effects, pp. 5162. Univ. Utah Press, Salt Lake City.

371

372 References
Mueggler, W. F. 1956. Is sagebrush seed residual in the soil of burns
or is it wind-borne? U.S. Forest Ser. Res. Note 35.
. 1975. Rate and pattern of vigor recovery in Idaho fescue and
bluebunch wheatgrass. J. Range Manage. 28:198204.
Muir, P. S., and J. E. Lotan. 1985a. Disturbance history and serotiny
of Pinus contorta in western Montana. Ecology 66:165868.
. 1985b. Serotiny and life history of Pinus contorta var lati
folia. Can. J. Bot. 63:93845.

National Academy of Sciences, Committee on Ungulate Management in Yellowstone National Park, National Research Council.
2002. Ecological Dynamics on Yellowstones Northern Range.
National Academy Press, Washington, D.C.
Naugle, D. E. 2011. Energy Development and Wildlife Conservation in Western North America. Island Press, Washington,
D.C.
Naugle, D. E., C. L. Aldridge, B. L. Walker, et al. 2004. West Nile

Mumma, S. A., C. Whitlock, and K. Pierce. 2012. A 28,000 year

virus: Pending crisis for greater sage-grouse. Ecol. Lett. 7:70413.

history of vegetation and climate from Lower Red Rock Lake,

. 2005. West Nile virus and sage-grouse: What more have we

Centennial Valley, southwestern Montana, USA. Palaeogeogr.


Palaeocl. 32628: 3041.
Mummey, D. L., P. E. Stahl, and J. S. Buyer. 2002. Soil microbiologi-

learned? Wildlife Soc. Bull. 33:61623.


Neff, D. J. 1957. Ecologial effects of beaver habitat abandonment in
the Colorado Rockies. J. Wildlife Manage. 21:8084.

cal properties 20 years after surface mine reclamation: Spatial

Negron, J. F., K. Allen, B. Cook, and J. R. Withrow. 2008. Suscepti-

analysis of reclaimed and undisturbed sites. Soil Biol. Biochem.

bility of ponderosa pine, Pinus ponderosa (Dougl. Ex Laws.), to

34:171725.

mountain pine beetle, Dendroctonus ponderosae Hopkins, attack

Munroe, J. S. 2003. Estimates of Little Ice Age climate inferred


through historical rephotography, northern Uinta Mountains,
USA. Arct. Antarct. Alp. Res. 35:48998.
. 2012. Physical, chemical, and thermal properties of soils
across a forest-meadow ecotone in the Uinta Mountains, northeastern Utah, USA. Arct. Antarct. Alp. Res. 44:95106.
Murray, R. B. 1975. Effect of Artemisia tridentata removal on mineral
cycling. Ph.D. diss., Washington State Univ., Pullman.
Muscha, J. M., and A. L. Hild. 2006. Biological soil crusts in grazed
and ungrazed Wyoming sagebrush steppe. J. Arid Environ.
67:195207.
Myers, W. E. 1969. Relationship of vegetation to geologic formations along Sheep Mountain, Albany County, Wyoming. M.S.
thesis, Univ. Wyoming, Laramie.
Nacify, C., A. Sala, E. G. Keeling, et al. 2010. Interactive effects of
historical logging and fire exclusion on ponderosa pine forest
structure in the northern Rockies. Ecol. Appl. 20:185164.
Naftz, D. L., D. D. Susong, P. F. Schuster, et al. 2002. Ice core evidence of rapid air temperature increases since 1960 in alpine
areas of the Wind River Range, Wyoming, United States. J. Geophys. Res. Atmosph. 107:ACL 3-1ACL 3-16.
Naftz, D. L., L. Oswald, P. F. Schuster, and K. Miller. 2009. Ice-core
and flood-flow evidence of rapid climate change, Fitzpatrick Wil-

in uneven-aged stands in the Black Hills of South Dakota and


Wyoming USA. Forest Ecol. Manage. 254:32734.
Neilson, R. P., and L. H. Wullstein. 1983. Biogeography of two southwestern American oaks in relation to seedling drought response
and atmospheric flow structure. Biogeography 10:27597.
Neilson, R. P., J. M. Lenihan, D. Bachelet, and R. J. Drapek. 2005.
The sage-grouse dilemma: A case study of long-term landscape
use and abuse. In Transactions of the 70th North American
Wildlife and Natural Resources Conference, pp. 14559. Wildlife Mgmt. Inst., Washington, D.C.
Nelson, D. L., and D. L. Sturges. 1986. A snowmold disease of
mountain big sagebrush. Phytopathology 76:94651.
Nelson, D. L., and C. F. Tiernan. 1983. Winter injury of sagebrush
and other wildland shrubs in the western United States. U.S.
Forest Ser. Res. Paper INT-314.
Nelson, K. N., M. E. Rocca, M. Diskin, et al. 2014. Predictors of
bark beetle activity and scale-dependent spatial heterogeneity
change during the course of an outbreak in a subalpine forest.
Landscape Ecol. 29:97109.
Nelson, N. A., and J. Pierce. 2010. Late-Holocene relationships
among fire, climate and vegetation in a forest-sagebrush ecotone of southwestern Idaho, USA. Holocene 20:117994.
Newbauer, J. J., III, L. M. White, R. M. Moy, and D. A. Perry. 1980.

derness Area, Wind River Range, Wyoming. In F. H. Wagner, ed.,

Effects of increased rainfall on native forage production in east-

Climate Warming in Western North America: Evidence and En-

ern Montana. J. Range Manage. 33:24650.

vironmental Effects, pp. 3548. Univ. Utah Press, Salt Lake City.

Newton, H., and W. P. Jenney. 1880. Report on the Geology and

Nagler, P. L., E. P. Glenn, C. S. Jarnevich, and P. B. Shafroth. 2011.

Resources of the Black Hills of Dakota, with Atlas. U.S. Gov.

Distribution and abundance of saltcedar and Russian olive in


the Western United States. Crit. Rev. Plant Sci. 30:50823.
Naithani, K. J., B. E. Ewers, and E. Pendall. 2012. Sap flux-scaled
transpiration and stomatal conductance response to soil and
atmospheric drought in a semi-arid sagebrush ecosystem.
J.Hydrol. 46465:17685.
Nsholm, T., A. Ekblad, A. Nordin, et al. 1998. Boreal forest plants
take up organic nitrogen. Nature 392:91416.
National Academy of Sciences, Committee on Mitigating Wetland
Losses, National Research Council. 2001. Compensating for
Wetland Losses under the Clean Water Act. National Academy
Press, Washington, D.C.

Printing Office, Washington, D.C.


Nichols, J. T. 1964. Soil-vegetation relationships of the fifteen-mile
drainage, Washakie County, Wyoming. Ph.D. diss., Univ. Wyoming, Laramie.
Nie, M., E. Pendall, C. Bell, et al. 2012. Positive climate feedbacks of
soil microbial communities in a semi-arid grassland. Ecol. Lett.
doi: 10.1111/ele.12034.
Nippert, J. B., and A. K. Knapp. 2007. Linking water uptake with
rooting patterns in grassland species. Oecologia. doi: 10.1007/
s00442-007-0745-8.
Norris, J. R. 2006. Influence of climate on the modern and late
Holo
cene biogeography of ponderosa pine in the central

References
Rockies. Ph.D. dissertation, Department of Botany, Univ. Wyoming, Laramie.
Norris, J. R., S. T. Jackson, and J. L. Betancourt. 2006. Classification

. 1982. Beating the walnut tree: More on grass/grazer mutualism. Oikos 39:11516.
Owen, S. M., C. H. Sieg, and C. A. Gehring. 2011. Rehabilitating

tree and minimum-volume ellipsoid analyses of the distribution

downy brome (Bromus tectorum)invaded shrublands using

of ponderosa pine in the western USA. J. Biogeogr. 3:34260.

Imazapic and seeding with native shrubs. Invasive Plant Sci.

Norton, J. B., T. A. Monaco, J. M. Norton, et al. 2004. Soil morphology and organic matter dynamics under Cheatgrass and sagebrush-steppe plant communities. J. Arid Environ. 57:44566.
Oberbauer, S. F., and W. D. Billings. 1981. Drought tolerance and
water use by plants along an alpine topographic gradient. Oecologia 50:32531.
Ode, D. J., and L. Tieszen. 1980. The seasonal contribution of C3
and C4 plant species to primary production in a mixed prairie.
Ecology 61:130411.
Oesterheld, M., J. Loreti, M. Semmartin, and J. M. Paruelo. 1999.
Grazing, fire, and climate effects on primary productivity of
grasslands and savannas. In L. R. Walker, ed., Ecosystems of
the World 16: Ecosystems of Disturbed Ground, pp. 287306.
Elsevier, Amsterdam.
Olliff, T., G. Plumb, J. Kershner, et al. 2010. A science agenda for
the Greater Yellowstone Area: Responding to landscape impacts
from climate change, land use change, and invasive species.
Yellowstone Sci. 18:1422.
Olmsted, C. E. 1979. The ecology of aspen with reference to utilization by large herbivores in Rocky Mountain National Park. In
M. S. Boyce and L. D. Hayden-Wing, eds., North American Elk:
Ecology, Behavior and Management, pp. 8997. Univ. Wyoming
Press, Laramie.
Olson, R. A., and W. A. Gerhart. 1982. A physical and biological
characterization of riparian habitat and its importance in Wyoming. Wyoming Game and Fish Dept., Cheyenne.
Oosting, H. J., and J. F. Reed. 1952. Virgin spruce-fir of the Medicine
Bow Mountains, Wyoming. Ecol. Monogr. 22:6991.

Manage. 4:22333.
Owens, M. K., and G. W. Moore. 2007. Saltcedar water use: Realistic and unrealistic expectations. Rangeland Ecol. Manage.
60:55357.
Painter, E. L., and J. Belsky. 1993. Application of herbivore optimization theory to rangelands of the western United States. Ecol.
Appl. 3:29.
Painter, E. L., and J. K. Detling. 1981. Effects of defoliation on net
photosynthesis and regrowth of western wheatgrass. J. Range
Manage. 34:6871.
Palmer, T. 1991. The Snake River: Window to the West. Island Press,
Covelo, California.
Parchman, T. L., Z. Gompert, J. Mudge, et al. 2012. Genome-wide
association genetics of an adaptive trait in lodgepole pine. Mol.
Ecol. 21:29913005.
Parker, A. J. 1986. Persistence of lodgepole pine forests in the central Sierra Nevada. Ecology 67:156067.
Parker, A. J., and K. C. Parker. 1983. Comparative successional roles
of trembling aspen and lodgepole pine in the southern Rocky
Mountains. Great Basin Nat. 43:44755.
Parker, J. 1953. Photosynthesis in Picea excelsa in winter. Ecology
34:6059.
Parkman, F. 1950. The Oregon Trail 1849. New American Library,
New York.
Parmenter, A. W., A. Hansen, R. E. Kennedy, et al. 2003. Land use
and land cover change in the Greater Yellowstone Ecosystem:
19751995. Ecol. Appl. 13:687703.
Parmenter, R. R., M. R. Mesch, and J. A. MacMahon. 1987. Shrub

Orabona, A., S. Patla, L. Van Fleet, et al. 2009. Atlas of Birds, Mam-

litter production in a sagebrush-steppe ecosystem: Rodent

mals, Amphibians, and Reptiles in Wyoming. Wyoming Game

population cycles as a regulating factor. J. Range Manage.

and Fish Department Nongame Program, Lander. http://wgfd


.wyo.gov/web2011/Departments/Wildlife/pdfs/WILDLIFE _
ANIMALATLAS0002711.pdf.
Orr, H. K. 1975. Watershed management in the Black Hills: The status of our knowledge. U.S. Forest Ser. Res Paper RM-141.
Ostlund, E. 2012. Dangerous Crossing. Natl. Geographic Daily
News, December 4, 2012.

40:5054.
Parsons, W. F. J., D. H. Knight, and S. L. Miller. 1994. Root gap dynamics in lodgepole pine forests: Nitrogen transformations in
gaps of different size. Ecol. Appl. 4:35462.
Passey, H. B., and V. K. Hugie. 1963. Variation in bluebunch wheatgrass in relation to environment and geographic location. Ecology 44:15861.

Ostoja, S. M., and E. W. Schupp. 2009. Conversion of sagebrush

Pataki, D. E., R. Oren, and W. K. Smith. 2000. Sap flux of co-

shrublands to exotic annual grasslands negatively impacts

occurring species in a western subalpine forest during seasonal

small mammal communities. Divers. Distrib. 15:86370.

soil drought. Ecology 81:255766.

Ostresh, L. M., R. A. Marston, and W. M. Hudson, eds. 1990. Wyo-

Patten, D. T. 1959. An ecological study of the parks of the Rocky

ming Water Atlas. Wyoming Water Development Commission,

Mountains in the Gallatin area of Montana. M.S. thesis, Univ.

Cheyenne.
Oswald, E. T. 1966. A synecological study of the forested moraines
of the valley floor of Grand Teton National Park, Wyoming.
Ph.D. diss., Montana State Univ., Bozeman.
Owen, D. F., and R. G. Wiegert. 1976. Do consumers maximize
plant fitness? Oikos 27:48892.
. 1981. Mutualism between grasses and grazers: An evolutionary hypothesis. Oikos 36:37678.

Massachusetts, Amherst.
. 1963. Vegetational patterns in relation to environments in
the Madison Range, Montana. Ecol. Monogr. 33:375406.
. 1968. Dynamics of the shrub continuum along the Gallatin
River in Yellowstone National Park. Ecology 49:110712.
Patten, R. S., and D. H. Knight. 1994. Snow avalanches and vegetation pattern in Cascade Canyon, Grand Teton National Park,
Wyoming, U.S.A. Arc. Alp. Res. 26:3541.

373

374 References
Paul, E. A., F. E. Clark, and V. O. Biederbeck. 1979. Micro-organisms.

Peters, D. P. C., W. K. Lauenroth, and I. C. Burke. 2008. The role

In R. T. Coupland. ed., Grassland Ecosystems of the World:

of disturbances in shortgrass steppe community and ecosys-

Analysis of Grasslands and Their Uses, pp. 8796. Cambridge

tem dynamics. In W. K. Lauenroth and I. C. Burke, eds., Ecol-

Univ. Press, Cambridge.

ogy of the Shortgrass Steppe, pp. 84118. Oxford Univ. Press,

Pauli, J. N., S. W. Buskirk, E. S. Williams, and W. H. Edwards. 2006.


A plague epizootic in the black-tailed prairie dog (Cynomys
ludovicianus). J. Wildlife Dis. 42:7480.
Pearce, C. M., and D. G. Smith. 2007. Invasive saltcedar (Tamarix):
Its spread from the American southwest to the northern Great
Plains. Phys. Geogr. 28:50730.

Oxford.
Petit, J. R., J. Jouzel, D. Raynaud, et al. 1999. Climate and atmospheric history of the past 420,000 years from the Vostok ice
core, Antarctica. Nature 399:42936.
Pw, T. L. 1983. Alpine permafrost in the contiguous United
States: A review. Arct. Alp. Res. 15:14556.

Pearson, J. A., D. H. Knight, and T. J. Fahey. 1987. Biomass and nu-

Pfadt, R. E., and D. M. Hardy. 1987. A historical look at rangeland

trient accumulation during stand development in Wyoming

grasshoppers and the value of grasshopper control programs. In

lodgepole pine forests. Ecology 68:196673.

J. L. Capinera, ed., Integrated Pest Management on Rangeland:

Pearson, L. C. 1965a. Primary production in grazed and ungrazed


desert communities of eastern Colorado. Ecology 46:27885.
. 1965b. Primary productivity in a northern desert area.
Oikos 15:21128.
Peck, D. E., and J. R. Lovvorn. 2001. The importance of flood irrigation in water supply to wetlands in the Laramie Basin, Wyoming, USA. Wetlands 21:37078.
Peck, D. E., D. M. McLeod, J. P. Hewlett, and J. R. Lovvorn. 2004.
Irrigation-dependent wetlands versus instream flow enhancement: Economics of water transfers from agriculture to wildfire
uses. Environ. Manage. 34:84255.
Peden, D. G. 1976. Botanical composition of bison diets on shortgrass plains. Am. Midl. Nat. 96:22529.
Pederson, G. T., L. J. Graumlich, D. B. Fagre, et al. 2010. A century

A Shortgrass Prairie Perspective, pp. 18395. Westview Press,


Boulder.
Pfeifer, E. M., J. A. Hicke, and A. J. H. Meddens. 2011. Observations
and modeling of aboveground tree carbon stocks and fluxes
following a bark beetle outbreak in the western United States.
Global Change Biol. 17:33950.
Pfister, J. A., F. D. Provenza, K. E. Panter, et al. 2002. Risk management to reduce livestock losses from toxic plants. J. Range Manage. 55:291300.
Philipps, D. 2012. Nowhere to run. High Country News (Paonia,
Colorado), November, pp.1220.
Phillips, C. M. 1977. Willow carrs of the upper Laramie River Valley,
Colorado. M.S. thesis, Colorado State Univ., Fort Collins.
Phillips, V. D. 1982. Responses by alpine plants and soils to micro-

of climate and ecosystem change in western Montana: What do

topography within sorted polygons. Ph.D. diss., Univ. Colo-

temperature trends portend? Climatic Change 98:13354.

rado, Boulder.

Pederson, G. T., S. T. Gray, T. Ault, et al. 2011. Climatic controls

Pielke, R. A., Sr., and N. J. Doesken. 2008. Climate of the shortgrass

on the snowmelt hydrology of the Northern Rocky Mountains.

steppe. In W. K. Lauenroth and I. C. Burke, eds., Ecology of the

J.Climate 24:166687.

Shortgrass Steppe, pp. 1429. Oxford Univ. Press, Oxford.

Peet, R. K. 2000. Forests and meadows of the Rocky Mountains. In


M. G. Barbour and W. D. Billings, eds., North American Terrestrial Vegetation, 2nd edition, pp. 75122. Cambridge Univ.
Press, Cambridge.
Pelz, K. A., and F. W. Smith. 2012. Thirty year change in lodgepole
and lodgepole/mixed conifer forest structure following 1980s
mountain pine beetle outbreak in western Colorado, USA. Forest Ecol. Manage. 280:93102.

Pierce, K. L. 2003. Pleistocene glaciations in the Rocky Mountains.


Devel. Quaternary Sci. 1:6376.
Platt, P., and N. Slater. 1852. Travelers Guide across the Plains
upon the Overland Route to California. John Howell Book, San
Francisco.
Platts, W. S. 1981. Effects of sheep grazing on a riparian stream
environment. U.S. Forest Ser. Res. Note INT-307.
. 1982. Livestock and riparian-fishery interactions: What are

Perera, A. H., L. J. Buse, and M. G. Weber. 2004. Emulating Natural

the facts? In K. Sabol, ed., Transactions of the North American

Forest Landscape Disturbances: Concepts and Applications. Co-

Wildlife Natural Resources Conference 47, pp. 50715. Wildlife

lumbia Univ. Press, New York.

Management Institute, Washington, D.C.

Perrings, C., S. Naeem, F. S. Ahrenstani, et al. 2011. Ecosystem ser-

Platts, W. S., W. F. Megahan, and G. W. Minshall. 1983. Methods for

vices, targets, and indicators for the conservation and sustain-

evaluating stream, riparian, and biotic conditions. U.S. Forest

able use of biodiversity. Front. Ecol. Environ. 9:51220.

Ser. Gen. Tech. Rep. INT-138.

Perry, L. G., D. M. Blumenthal, T. A. Monaco, et al. 2010. Im-

Plummer, A. P., D. R. Christensen, S. B. Monsen. 1968. Restoring

mobilizing nitrogen to control plant invasion. Oecologia 163:

big game range in Utah. Publication 68-3. Utah Div. Fish Game,

1324.

Salt Lake City, Utah.

Perry, T. O. 1971. Winter-season photosynthesis and respiration by

Pocewicz, A., J. M. Kiesecker, G. P. Jones, et al. 2011. Effectiveness

twigs and seedlings of deciduous and evergreen trees. Forest Sci.

of conservation easements for reducing development and main-

17:4143.

taining biodiversity in sagebrush ecosystems. Biol. Conserv.

Persico, L., and G. Meyer. 2009. Holocene beaver damming, fluvial geomorphology, and climate in Yellowstone National Park,
Wyoming. Quaternary Res. 71:34053.

144:56774.
Pocewicz, A., W. A. Estes-Zumpf, M. D. Andersen, et al. 2013.
Modeling the distribution of migratory bird stopovers to

References
inform landscape-scale siting of wind development. PLoS ONE
8:e75363. doi: 10.1371/journal.pone.0075363.
Pocknall, D. T., and R. M. Flores. 1987. Coal palynology and sedimentology in the Tongue River Member, Fort Union Formation,
Powder River Basin, Wyoming. Palaios 2:13345.
Poff, B., K. A. Koestner, D. G. Neary, and V. Henderson. 2011. Threats
to riparian ecosystems in Western North America: An analysis
of existing literature. J. Am. Water Resour. Assoc. 47:124154.
Poiani, K. A., and W. C. Johnson. 1991. Global warming and prairie
wetlands. BioScience 41:61118.
. 1993. Potential effects of climate change on a semipermanent prairie wetland. Climatic Change 24:21332.
Polley, H. W., and J. K. Detling. 1988. Herbivory tolerance of Agropyron smithii populations with different grazing histories. Oecologia 77:26167.
Poore, R. E., C. A. Lamanna, J. J. Ebersole, and B. J. Enquist. 2009.
Controls on radial growth of mountain big sagebrush and implications for climate change. West. Midl. Natural. 69:55662.
Porter, E., and S. Johnson. 2007. Translating science into policy:
Using ecosystem thresholds to protect resources in Rocky
Mountain National Park. Environ. Pollution 149:26880.
Potter, L. D., and D. L. Green. 1964. Ecology of ponderosa pine in
western North Dakota. Ecology 45:1023.
Potvin, M. A., and A. T. Harrison. 1984. Vegetation and litter
changes of a Nebraska sandhills prairie protected from grazing.
J. Range Manage. 37:5558.
Powell, E. N., P. A. Townsend, and K. F. Raffa. 2012. Wildfire provides refuge from local extinction but is an unlikely driver of
outbreaks by mountain pine beetle. Ecol. Monogr. 82:6984.
Power, T. M. 2005. Ecosystem preservation and the economy in the
Greater Yellowstone area. Conserv. Biol. 5:395404.
Prasciunas, M. M., G. C. Frison, M. Kornfeld, et. al. 2008. Clovis in
Wyoming. Curr. Res. Pleistocene 25:13538.
Prater, M. R., D. Obrist, J. A. Arnone III, and E. H. DeLucia. 2006.
Net carbon exchange and evapotranspiration in postfire and
intact sagebrush communities in the Great Basin. Oecologia
146:595607.
Preisler, H. K., J. A. Hicke, A. A. Ager, and J. L. Hayes. 2012. Climate
and weather influences on spatial temporal patterns of mountain pine beetle populations in Washington and Oregon. Ecology 93:242134.
Prescott, C. E., J. P. Corbin, and D. Parkinson. 1989. Biomass, productivity and nutrient-use efficiency of aboveground vegetation in four Rocky Mountain coniferous forests. Can. J. Forest
Res. 19:30917, 48998.
Preuss, C. 1958. Exploring with Fremont. E. G. Gudde and E. K.
Gudde, eds. and trans. Univ. Oklahoma Press, Norman.

Prieto, I. F., M. Padilla, C. Armas, and F. I. Pugnaire. 2011. The


role of hydraulic lift on seedling establishment under a nurse
plant species in a semi-arid environment. Perspect. Plant Ecol.
13:18187.
Pritchard, J. A. 1999. Preserving Yellowstones Natural Conditions.
Univ. Nebraska Press, Lincoln.
Progulske, D. R. 1974. Yellow ore, yellow hair, yellow pine: A photographic study of a century of forest ecology. South Dakota Agri.
Exp. Sta. Bull. 616:1169.
Proulx, A., ed. 2008. Red Desert: History of a Place. Univ. Texas
Press, Austin.
Pruett, C. L., M. A. Patten, and D. H. Wolfe. 2009. Its not easy being
green: Wind energy and a declining grassland bird. BioScience
59:25762.
Pugh, E., and E. Gordon. 2012. A conceptual model of water yield
effects from beetle-induced tree death in snow-dominated
lodgepole pine forests. Hydrol. Process. doi: 10.1002/hyp.9312.
Pugh, E., and E. Small. 2012. The impact of pine beetle infestation
on snow accumulation and melt in the headwaters of the Colorado River. Ecohydrology 5:46777.
Quinn, M. A., and D. D. Walgenbach. 1990. Influence of grazing history on the community structure of grasshoppers of a
mixed-grass prairie. Environ. Ent. 19:175666.
Raffa, K. F., B. H. Aukema, B. J. Bentz, et al. 2008. Cross-scale drivers
of natural disturbances prone to anthropogenic amplification:
Dynamics of biome-wide bark beetle eruptions. BioScience 58:
50117.
Randall, D. 1983. A new role for beaver. Defenders 58:2932.
Rau, B. M., D. W. Johnson, R. R. Blank, et al. 2011. Transition from
sagebrush steppe to annual grass (Bromus tectorum): Influence
on belowground carbon and nitrogen. Rangeland Ecol. Manage. 64:13947.
Rauzi, F., and M. L. Fairbourn. 1983. Effects of annual applications
of low N fertilizer rates on a mixed grass prairie. J. Range Manage. 36:35962.
Rauzi, F., and F. M. Smith. 1973. Infiltration rates: Three soils with
three grazing levels in northeastern Colorado. J. Range Manage.
26:12629.
Raventon, E. 1994. Island in the Plains: A Black Hills Natural History. Johnson Books, Boulder.
Raynolds, W. F. 1868. Report of Brevet Colonel W. F. Raynolds,
U.S.A., Corps of Engineers, on the exploration of the Yellowstone and Missouri rivers, in 185960. 40th Congress. Serial
1317. Senate Exec. Doc. 77.
Redman, R. E. 1978. Plant and soil water potentials following fire
in a northern mixed grassland. J. Range Manage. 31:44345.

Prevey, J. S., M. J. Germino, and N. J. Huntly. 2010a. Loss of foun-

Reed, M. J., and R. A. Peterson. 1961. Vegetation, soil, and cattle

dation species increases population growth of exotic forbs in

responses to grazing on northern Great Plains range. U.S. Forest

sagebrush steppe. Ecol. Appl. 20:18901902.


Prevey, J. S., M. J. Germino, N. J. Huntly, and R. S. Inouye. 2010b.
Exotic plants increase and native plants decrease with loss of
foundation species in sagebrush steppe. Plant Ecol. 207:3951.
Price, M. V., and N. M. Waser. 2000. Responses of subalpine
meadow vegetation to four years of experimental warming.
Ecol. Appl. 10:81123.

Ser. Tech. Bull. 1252.


Reed, R. M. 1971. Aspen forests of the Wind River Mountains, Wyoming. Am. Midl. Nat. 86:32743.
. 1976. Coniferous forest habitat types of the Wind River
Mountains, Wyoming. Am. Midl. Nat. 95:15973.
Reese, K. P., and J. W. Connelly. 2011. Harvest management for
greater sage-grouse: A changing paradigm for game bird

375

376 References
management. In S. T. Knick and J. W. Connelly, eds., Greater

Rhoades, C. C., D. Entwistle, and D. Butler. 2011. The influence

Sage-grouse: Ecology and Conservation of a Landscape Species

of wildfire extent and severity on streamwater chemistry, sedi-

and Its Habitats, pp. 10111. Studies in Avian Biology 38. Univ.

ment and temperature following the Hayman Fire, Colorado.

California Press, Berkeley.

Int. J. Wildlife Fire 20:43042.

Regniere, J., and B. Bentz. 2007. Modeling cold tolerance in the

Rhoades, C. C., J. H. McCutchan, Jr., L. A. Cooper, et al. 2013. Bio

mountain pine beetle, Dendroctonus ponderosae. J. Insect Physiol.

geochemistry of beetle-killed forests: Explaining a weak ni-

53:55972.

trate response. Proc. Natl. Acad. Sci. USA. doi: 10-1073/pnas.

Reher, C. A., and G. C. Frison. 1980. The Vore Site: A stratified buffalo jump in the Wyoming Black Hills. Plains Anthropologist
Memoir 16 (vol. 25).
Rehfeldt, G. E. 1985. Ecological genetics of Pinus contorta in the
Wasatch and Uinta Mountains of Utah. Can. J. Forest Res. 15:
52430.
Rehfeldt, G. E., N. L. Crookston, M. V. Warwell, and J. S. Evans.
2006. Empirical analyses of plant-climate relationships for the
western United States. Int. J. Plant Sci. 167:112350.
Reichardt, K. L. 1982. Succession of abandoned fields on the
shortgrass prairie, northeastern Colorado. Southwest. Nat. 27:
299304.
Reider, R. G. 1983. A soil catena in the Medicine Bow Mountains,
Wyoming, USA, with reference to paleoenvironmental influences. Arct. Alp. Res. 15:18192.
. 1990. Late Pleistocene and Holocene pedogenic and environmental trends at archaeological sites in plains and mountain
areas of Colorado and Wyoming. In N. P. Lasca and J.Donahue,
eds., Archaeological Geology of North America, pp. 33560.
Centennial Special Vol. 4. Geol. Soc. Am., Boulder.
Reider, R. G., G. A. Huckleberry, and G. C. Frison. 1988. Soil evidence for postglacial forest-grassland fluctuation in the Absaroka Mountains of northwestern Wyoming, U.S.A. Arct. Alp.
Res. 20:18898.
Reider, R. G., J. M. Huss, and T. W. Miller. 1996. A groundwater
vortex hypothesis for mima-like mounds, Laramie Basin, Wyoming. Geomorphology 16:295317.
. 1999. Stratigraphy, soils, and age relationships of mima-like
mounds, Laramie Basin, Wyoming. Phys. Geogr. 20:8396.
Reiners, W. A. 2003. Natural ecosystems I. The Rocky Mountains,
In F. H. Wagner, ed. Rocky Mountain/Great Basin Regional Climate-Change Assessment. Report for the U.S. Global Change
Research Program, pp. 14584. Utah State Univ., Logan.
Renkin, R. A., and D. G. Despain. 1992. Fuel moisture, forest type,
and lightning-caused fire in Yellowstone National Park. Can. J.
Forest Res. 22:3745.
Rennick, R. B. 1981. Effects of prescribed burning on mixed prairie
vegetation in southeastern Montana. M.S. thesis, Montana
State Univ., Bozeman.
Resler, L. M., and D. F. Tomback. 2008. Blister rust prevalence in

1221029110.
Rice, J., A. Tredennick, and L. A. Joyce. 2012. Climate change on
the Shoshone National Forest, Wyoming: A synthesis of past
climate, climate projections, and ecosystem implications. U.S.
Forest Ser. Gen. Tech. Rep. RMRS-GTR-264.
Richards, J. H., and M. M. Caldwell. 1985. Soluble carbohydrates,
concurrent photosynthesis and efficiency in regrowth following defoliation: A field study with Agropyron species. J. Appl.
Ecol. 22:90720.
Richards, J. H., and L. A. Donovan. 2005. Multi-scale spatial
variation in hydraulic lift by the desert halophyte, Sarcobatus
vermiculatus: Implications for resource acquisition from dry
soils. Comp. Biochem. Phys. A 141:S317.
Richmond, G. M. 1949. Stone nets, stone stripes, and soil stripes in
the Wind River Mountains, Wyoming. J. Geol. 57:14353.
Rickard, W. H. 1964. Demise of sagebrush through soil changes.
BioScience 14:4344.
. 1967. Seasonal soil moisture patterns in adjacent greasewood and sagebrush stands. Ecology 48:103438.
Riedl, G. 1959. Geology of the eastern portion of Shirley Basin, Albany and Carbon counties, Wyoming. M.S. thesis, Univ. Wyoming, Laramie.
Righter, R. W. 1982. Crucible for Conservation: The Creation of
Grand Teton National Park. Colorado Associated Univ. Press,
Boulder.
Ripple, W. J., and R. L. Beschta. 2004. Wolves and the ecology
of fear: Can predation risk structure ecosystems? BioScience
54:75566.
Ripple, W. J., and E. J. Larsen. 2000. Historic aspen recruitment,
elk, and wolves in northern Yellowstone National Park, USA.
Biol. Conserv. 95:36170.
Risser, P. G. 1985. Grasslands. In B. F. Chabot and H. A. Mooney,
eds., Physiological Ecology of North American Plant Communities, pp. 23256. Chapman and Hall, New York.
Risser, P. G., and W. J. Parton. 1982. Ecosystem analysis of the tallgrass prairie: Nitrogen cycle. Ecology 63:134251.
Rissman, A. R., L. Lozier, T. Comendant, et al. 2007. Conservation
easements: Biodiversity protection and private use. Conserv.
Biol. 21:70918.

krummholz whitebark pine: Implications for treeline dynam-

Roach, J. L., P. Stapp, B. Van Horne, and M. F. Antolin. 2001.

ics, northern Rocky Mountains, Montana, USA. Arct. Antarc.

Genetic structure of a metapopulation of black-tailed prairie

Alp. Res. 40:16170.

dogs. J. Mammol. 82:94659.

Reynolds, D. N. 1984. Alpine annual plants: Phenology, germina-

Robertson, D. R., J. L. Nielsen, and J. H. Bare. 1966. Vegetation and

tion, photosynthesis, and growth of three Rocky Mountain

soils of alkali sagebrush and adjacent big sagebrush ranges in

species. Ecology 65:75966. See also Oecologia 62:25055.

North Park, Colorado. J. Range Manage. 19:1720.

Reynolds, L. V., and D. J. Cooper. 2011. Ecosystem response to re-

Robertson, D. S., M. C. McKenna, O. B. Toon, et al. 2004. Sur-

moval of exotic riparian shrubs and a transition to upland veg-

vival in the first hours of the Cenozoic. Geol. Soc. Am. Bull.

etation. Plant Ecol. 212:124361.

116:76068.

References
Robertson, J. H. 1947. Response of range grasses to different intensities of competition with sagebrush. Ecology 28:116.
Rochow, T. F. 1970. Ecological investigations of Thlaspi alpestre L.
along an ecological gradient in the central Rocky Mountains.
Ecology 51:64956.
Rodel, C. F. 1977. A grasshopper model for a grassland ecosystem.
Ecology 58:22745.
Roe, F. E. 1951 The North American Buffalo: A Critical Study of the
Species in Its Wild State. Univ. Toronto Press, Toronto.

Rowe, C. L. J., and E. A. Leger. 2011. Competitive seedlings and


inherited traits: A test of rapid evolution of Elymus multisetus
(big squirreltail) in response to cheatgrass invasion. Evol. Appl.
4:48598.
Running, S. W. 1980. Environmental and physiological control of
water flux through Pinus contorta. Can. J. Forest Res. 10:8291.
Russey, G. R. 1967. The effect of grazing intensities on the roots
of Nuttalls saltbush (Atriplex nuttallii). M.S. thesis, Univ. Wyoming, Laramie.

Rogers, L. E. 1987. Ecology and management of harvester ants in

Rutherford, W. H. 1954. Interrelationships of beavers and other

the short grass plains. In J. L. Capinera, ed., Integrated Pest

wildlife on a high-altitude stream in Colorado. M.S. thesis,

Management on Rangeland, pp. 26170. Westview Press,


Boulder.
Romme, W. H. 1982. Fire and landscape diversity in subalpine forests of Yellowstone National Park. Ecol. Monogr. 52:199221.
Romme, W. H., and D. G. Despain. 1989. Historical perspective on
the Yellowstone fires of 1988. BioScience 39:69599.
Romme, W. H., and D. H. Knight. 1981. Fire frequency and subalpine forest succession along a topographic gradient in Wyoming. Ecology 62:31926.
. 1982. Landscape diversity: The concept applied to Yellowstone Park. BioScience 32:66470.
Romme, W. H., and M. G. Turner. 1991. Implications of global climate change for biogeographic patterns in the Greater Yellowstone Ecosystem. Conserv. Biol. 5:37386.
Romme, W. H., D. H. Knight, and J. B. Yavitt. 1986. Mountain pine
beetle outbreaks in the Rocky Mountains: Regulators of primary productivity? Am. Nat. 127:48494.
Romme, W. H., M. G. Turner, L. L. Wallace, and J. S. Walker. 1995.
Aspen, elk, and fire in northern Yellowstone National Park.
Ecology 76:20972106.
Romme, W. H., M. G. Turner, R. H. Gardner, et al. 1997. A rare
episode of sexual reproduction in Aspen (Populus tremulo
ides Michx) following the 1988 Yellowstone fires. Nat. Areas J.
17:1725.
Romme, W. H., M. G. Turner, G. A. Tuskan, and R. A. Reed. 2005.
Establishment, persistence, and growth of aspen (Populus tremuloides) seedlings in Yellowstone National Park. Ecology 86:
40418.
Romme, W. H., C. D. Allen, J. D. Balley, et al. 2009. Historical and
modern disturbance regimes, stand structures, and landscape
dynamics in pinyon-juniper vegetation of the western United
States. Rangeland Ecol. Manage. 62:20322.
Romme, W. H., M. S. Boyce, R. Gresswell, et al. 2011. Twenty years
after the 1988 Yellowstone fires: Lessons about disturbance and
ecosystems. Ecosystems 14:11961215.
Romo, J. T. 1984. Water relations in Artemisia tridentata subsp.
wyoming
ensis, Sarcobatus vermiculatus, and Kochia prostrata.
Ph.D. diss., Oregon State Univ., Corvallis.

Colorado State Univ., Fort Collins.


Ryan, M., and M. Squillace. 2004. Laws and regulations pertaining to wetland areas in the Intermountain West. In M. C. Mc
Kinstry, W. A. Hubert, and S. H. Anderson, eds., Wetland and
Riparian Areas of the Intermountain West: Ecology and Management, pp. 123. Univ. Texas Press, Austin.
Ryan, M. G., D. Binkley, J. H. Fownes, et al. 2004. An experimental
test of the causes of forest growth decline with stand age. Ecol.
Monogr. 74:393414.
Rydin, H., and J. Jeglum. 2006. The Biology of Peatlands. Oxford
Univ. Press, Oxford.
Ryel, R. J., M. M. Caldwell, C. K. Yoder, et al. 2002. Hydraulic redistribution in a stand of Artemisia tridentata: Evaluation of
benefits to transpiration assessed with a simulation model.
Oecologia 130:17384.
Ryel, R. J., A. J. Leffler, M. S. Peek, et al. 2004. Water conservation
in Artemisia tridentata through redistribution of precipitation.
Oecologia 141:33545.
Saab, V. A., Q. S. Latif, M. M. Rowland, et al. 2014. Ecological consequences of mountain pine beetle outbreaks for wildlife in western North American forests. Forest Sci. 60 [in press].
Sabinske, D. W., and D. H. Knight. 1978. Variation within the sagebrush vegetation of Grand Teton National Park. Northw. Sci.
52:195204.
Sacket, D. B. 1877. An intinerary of the route marched over on a reconnaissance along and through the Bighorn Mountains. In Report of Inspection by Generals P. H. Sheridan and W. T. Sherman,
pp. 611. U.S. Gov. Printing Office, Washington, D.C.
Sakai, A., and K. Otsuka. 1970. Freezing resistance of alpine plants.
Ecology 51:66571.
Sakai, A., and C. J. Weiser. 1973. Freezing resistance of trees in North
America with reference to tree regions. Ecology 54:11826.
Sala, A., S. D. Smith, and D. A. Devitt. 1996. Water use by Tamarix
ramosissima and associated phreatophytes in a Mohave desert
floodplain. Ecol. Appl. 6:88898.
Sala, O. E., W. J. Parton, L. A. Joyce, and W. K. Lauenroth. 1988. Primary production of the central grassland region of the United
States. Ecology 69:4045.
Salisbury, F. B. 1984. Light conditions and plant growth under the

Romo, J. T., and L. E. Eddleman. 1985. Germination response of

snow. In J. F. Merrit, ed., Winter Ecology of Small Mammals,

greasewood (Sarcobatus vermiculatus) to temperature, water po-

pp. 3950. Special Publication 10. Carnegie Mus. Nat. Hist.,

tential and specific ions. J. Range Manage. 38:11720.

Pittsburgh.

Rood, S. B., J. H. Braatne, and L. A. Goater. 2010. Favorable frag-

Sanderson, J. S., N. B. Kotliar, and D. A. Steingraeber. 2008. Oppos-

mentation: River reservoirs can impede downstream expansion

ing environmental gradients govern vegetation zonation in an

of riparian weeds. Ecol. Appl. 20:166467.

intermountain playa. Wetlands 28:106070.

377

378 References
Sauer, R. H. 1978. Effects of removal of standing dead material on

Park area prior to 1882. In J. D. Varley and W. G. Brewster, eds.,

growth of Agropyron spicatum. J. Range Manage. 31:12122.

Wolves in Yellowstone? A Report to the United States Congress.

Sawyer, H., R. M. Nielson, F. Lindzey, et al. 2007. Habitat selection


of Rocky Mountain elk in a nonforested environment. J. Wildlife Manage. 71:86874.
Sawyer, H., M. J. Kaufmann, and R. M. Nielson. 2009. The influence
of well pad activity on winter habitat selection patterns of mule
deer. J. Wildlife Manage. 73:105261.
Scheintaub, M. R., J. D. Derner, E. F. Kelly, and A. K. Knapp. 2009.
Response of the shortgrass steppe plant community to fire.
J.Arid Environ. 73:113643.
Schimpf, D. J., J. A. Henderson, and J. A. MacMahon. 1980. Some
aspects of succession in the spruce-fir forest zone of northern
Utah. Great Basin Nat. 40:126.

Vol. 4, Research and Analysis, pp. 1-31-174. National Park Service, Yellowstone National Park, Wyoming.
Schulte, P., L. Alegret, I. Arenillas, et al. 2010. The Chicxulub aster
oid impact and mass extinction at the Cretaceous-Paleogene
boundary. Science 327:121418.
Schulz, T. T., and W. C. Leininger. 1990. Differences in riparian
vegetation structures between grazed areas and exclosures.
J.Range Manage. 43:29599.
Schuman, G. E., L. E. Vicklund, and S. E. Belden. 2005. Establishing
Artemisia tridentata ssp. wyomingensis on mined lands: Science
and economics. Arid Land Res. Manage. 19:35362.
Schuster, W. S., D. L. Alles, and J. B. Mitton. 1989. Gene flow in

Schlaepfer, D. R., W. K. Lauenroth, and J. B. Bradford. 2011. Eco-

limber pine: Evidence from pollination phenology and ge-

hydrological niche of sagebrush ecosystems. Ecohydrology 5:

netic differentiation along an elevational transect. Am. J. Bot.

45366.

76:13951403.

Schmeisser, R. L., D. B. Loope, and D. A. Wedin. 2009. Clues to the

Schwartz, C. C., and J. Ellis. 1981. Feeding ecology and niche sepa-

medieval destabilization of the Nebraska Sand Hills, USA, from

ration in some native and domestic ungulates on the shortgrass

ancient pocket gopher burrows. Palaios 24:80917.

prairie. J. Appl. Ecol. 18:34353.

Schmid, J. M., and T. E. Hinds. 1974. Development of spruce-fir

Schwartz, C. C., M. A. Haroldson, G. C. White, et al. 2006. Tempo-

stands following spruce beetle outbreaks. U.S. Forest Ser. Res.

ral, spatial, and environmental influences on the demographics

Paper RM-131.

of grizzly bears in the Greater Yellowstone Ecosystem. Wildlife

Schoennagel, T., and C. R. Nelson. 2011. Restoration relevance of


recent National Fire Plan treatments in forests of the western
United States. Front. Ecol. Environ. 9:27177.
Schoennagel, T., M. G. Turner, and W. H. Romme. 2003. The influence of fire interval and serotiny on postfire lodgepole pine
density in Yellowstone National Park. Ecology 84:296778.

Monogr. 161:168.
Scott, D., and W. D. Billings. 1964. Effects of environmental factors
on standing crop and productivity of an alpine tundra. Ecol.
Monogr. 34:24370.
Scott, H. W. 1951. The geological work of the mound building ants
in western United States. J. Geol. 59:17375.

Schoennagel, T., T. T. Veblen, and W. H. Romme. 2004. The inter

Scott, J. W., S. E. Meyer, K. R. Merrill, and V. J. Anderson. 2010. Local

action of fire, fuels and climate across Rocky Mountain forests.

population differentiation in Bromus tectorum L. in relation to

BioScience 54:66176.

habitat-specific selection regimes. Evol. Ecol. 24:106180.

Schoennagel, T., R. L. Sherriff, and T. T. Veblen. 2011. Fire history

Scott, M. L., G. T. Auble, and J. M. Friedman. 1997. Flood depen-

and tree recruitment in the Colorado Front Range upper mon-

dency of cottonwood establishment along the Missouri River,

tane zone: Implications for forest restoration. Ecol. Appl. 21:


221022.
Schoennagel, T., T. T. Veblen, J. F. Negron, and J. M. Smith. 2012.
Effects of mountain pine beetle on fuels and expected fire behavior in lodgepole pine forests, Colorado, USA. PLoS ONE
7:e30002. doi: 10.1371/journal.pone.0030002.
Schoettle, A. W. 1991. The interaction between leaf longevity and

Montana, USA. Ecol. Appl. 7:67790.


Scott, R. L., W. L. Cable, and K. R. Hultine. 2008. The ecohydrologic
significance of hydraulic redistribution in a semiarid savanna.
Water Resour. Res. 44. doi: 10.1029/2007WR006149.
Scott, R. W. 1995. The Alpine Flora of the Rocky Mountains. Univ.
Utah Press, Salt Lake City.
Seager, S. T., C. Eisenberg, and S. B. St. Clair. 2013. Patterns and

shoot growth and foliar biomass per shoot in Pinus contorta at

consequences of ungulate herbivory on aspen in western North

two elevations. Tree Phys. 7:20914.

America. Forest Ecol. Manage. 299:8190.

Schreier, C. 1982. Grand Teton Explorers Guide. Homestead Publishing, Moose, Wyoming.
. 1983. Yellowstone Explorers Guide. Homestead Publishing,
Moose, Wyoming.
Schullery, P. 1989. The fires and fire policy. BioScience 39:68694.

Seastedt, T. R. 2001. Soils. In W. D. Bowman and T. R. Seastedt, eds.,


Structure and Function of an Alpine Ecosystem, pp. 15776. Oxford Univ. Press, New York.
Seastedt, T. R., and G. A. Adams. 2001. Effects of mobile tree islands
on alpine tundra soils. Ecology 82:817.

. 2004. Searching for Yellowstone: Ecology and Wonder in

Seastedt, T. R., W. D. Bowman, T. N. Caine, et al. 2004. The land-

the Last Wilderness, revised edition. Montana Historical Soci-

scape continuum: A model from high-elevation ecosystems.

ety Press, Helena.


. 2010. Greater Yellowstone science: Past, present, and future.
Yellowstone Sci. 18:713.

BioScience. 54:11121.
Sedgwick, J. A., and F. L. Knopf. 1989. Demography, regeneration,
and future projections for a bottomland cottonwood commu-

Schullery, P., and L. Whittlesey. 1992. The documentary record of

nity. In R. Sharitz and J. W. Gibbons, eds., Freshwater Wetlands

wolves and related wildlife species in the Yellowstone National

and Wildlife, pp. 24966. Conference Series 61. U.S. Dept.

References
Energy, Office of Scientific and Technical Information, Oak
Ridge, Tennessee.
. 1991. Prescribed grazing as a secondary impact in a western
riparian floodplain. J. Range Manage. 44:36973.
Seefeldt, S.S., M. Germino, and K. DiCristina. 2007. Prescribed fires
in Artemisia tridentata ssp. vaseyana steppe have minor and transient effects on vegetation cover and composition. Appl. Veg.
Sci. 10:24956.
Senft, R. L., L. R. Rittenhouse, and R. G. Woodmansee. 1985. Factors influencing patterns of cattle grazing behavior on shortgrass steppe. J. Range Manage. 38:8287.
Severson, K. E. 1963. A description and classification by composi-

Sherriff, R. L., and T. T. Veblen. 2006. Ecological effects of changes


in fire regimes in Pinus ponderosa ecosystems in the Colorado
Front Range. J. Veg. Sci. 17:70518.
. 2007. A spatially explict reconstruction of historical fire
occurrence in the ponderosa pine zone of the Colorado Front
Range. Ecosystems 10:31123.
Shinker, J. J., B. N. Shuman, T. A. Minckley, and A. K. Henderson.
2010. Climatic shifts in the availability of contested waters: A
long-term perspective from the headwaters of the North Platte
River. Ann. Assoc. Am. Geogr. 100:86679.
Shinneman, D. J., and W. L. Baker. 1997. Nonequilibrium dynamics between catastrophic disturbances and old-growth forests

tion of the aspen stands in the Sierra Madre Mountains, Wyo-

in ponderosa pine landscapes of the Black Hills. Conserv. Biol.

ming. Ph.D. diss., Univ. Wyoming, Laramie.

11:127688.

Severson, K. E., and J. J. Kranz. 1976. Understory production not

. 2009. Environmental and climatic variables as potential

predictable from aspen basal area or density. U.S. Forest Ser. Res.

drivers of post-fire cover of cheatgrass (Bromus tectorum) in

Note RM-314.

seeded and unseeded semiarid ecosystems. Int. J. Wildland Fire

Severson, K. E., and J. F. Thilenius. 1976. Classification of quaking


aspen stands in the Black Hills and Bear Lodge Mountains. U.S.
Forest Ser. Res. Paper RM-166.
Sexton, J. P., A. Sala, and K. Murray. 2006. Occurrence, persistence,
and expansion of saltcedar (Tamarix spp.) populations in the
Great Plains of Montana. West. N. Am. Nat. 66:111.
Shafroth, P. 2005. Tamarisk tensions. BioScience 55:821.
Shah, J. J. F., M. J. Herner, and T. M. Tibbets. 2010. Elaeagnus

18:191202.
Shinneman, D. J., W. L. Baker, and P. Lyon. 2008. Ecological restoration needs derived from reference conditions for a semi-arid
landscape in Western Colorado. J. Arid Environ. 72: 20727.
Shinneman, D. J., W. L. Baker, P. C. Rogers, and D. Kulakowski.
2013. Fire regimes of quaking aspen in the Mountain West. Forest Ecol. Manage. 299:2234.
Shuman, B. 2012. Recent Wyoming temperature trends, their driv-

angustifolia elevates soil inorganic nitrogen pools in riparian

ers, and impacts in a 14,000-year context. Climatic Change 112:

ecosystems. Ecosystems 13:4661.

42947.

Sharma, M. L., and D. J. Tongway. 1973. Plant-induced soil salinity

Shuman, B., A. Henderson, S. N. Coleman, et al. 2009. Holocene

patterns in two saltbush (Atriplex spp.) communities. J. Range

lake-level trends in the Rocky Mountains, USA. Quaternary Res.

Manage. 26:12125.
Sharp, L. A., K. Sanders, and N. Rimbey. 1990. Forty years of change
in a shadscale stand in Idaho. Rangelands 12:31328.
Sharpe, P. B., and B. Van Horne 1998. Influence of habitat on behavior of Townsends ground squirrels (Spermophilus townsendii).
J.Mammol. 79:90618.
Sharps, J. C., and D. W. Uresk. 1990. Ecological review of blacktailed prairie dogs and associated species in western South Dakota. Great Basin Nat. 50:33945.
Shaw, M. R., M. E. Loik, and J. Harte. 2000. Gas exchange and water
relations of two Rocky Mountain shrub species exposed to a
climate change manipulation. Plant Ecol. 146:197206.
Shea, K. 1985. Demographic aspects of coexistence in Engelmann
spruce and subalpine fir. Am. J. Bot. 72:182333.

Rev. 28:196179.
Shumar, M. L., and J. E. Anderson. 1986. Gradient analysis of vegetation dominated by two subspecies of big sagebrush. J. Range
Manage. 39:15660.
Shumway, S. W. 2000. Facilitative effects of a sand dune shrub
on species growing beneath the shrub canopy. Oecologia 124:
13848.
Sibold, J. S., T. T. Veblen, and M. E. Gonzales. 2006. Spatial and
temporal variation in historic fire regimes in subalpine forests
across the Colorado Front Range in Rocky Mountain National
Park, Colorado, USA. J. Biogeogr. 32:63147.
Sibul, J. R. 1995. Soil morphology and park dynamics of Dry Park,
Medicine Bow Mountains, Wyoming. M.S. thesis, Univ. Wyoming, Laramie.

Sheets, W. B. 1958. The effect of chemical control of big sagebrush

Siepielski, A. M., and C. W. Benkman. 2007. Convergent patterns

on the water budget and vertical energy balance. Ph.D. diss.,

in the selection mosaic for two North American bird-dispersed

Univ. Wyoming, Laramie.


Sheppard, J. S. 1971. The influence of geothermal temperature gradients upon vegetation patterns in Yellowstone National Park.
Ph.D. diss., Colorado State Univ., Fort Collins.
Shepperd, W. D., and M. A. Battaglia. 2002. Ecology, silviculture,
and management of Black Hills ponderosa pine. U.S. Forest Ser.
RMRS-GTR-97.

pines. Ecol. Monogr. 77:20320.


. 2008. Seed predation and selection exerted by a seed predator influence subalpine tree densities. Ecology 89:296066.
Sievering, H. 2001. Atmospheric chemistry and deposition. In
W.D. Bowman and T. R. Seastedt, eds., Structure and Function
of an Alpine Ecosystem, pp. 3244. Oxford Univ. Press, Oxford.
Simard, M., W. H. Romme, J. M. Griffin, and M. G. Turner. 2011. Do

Sher, A., and M. F. Quigley. 2013. Tamarix: A Case Study of Eco-

mountain pine beetle outbreaks change the probability of ac-

logical Change in the American West. Oxford Univ. Press,

tive crown fire in lodgepole pine forests? Ecol. Monogr. 81:324

Oxford.

and Ecology 93:94146.

379

380 References
Simard, M., E. N. Powell, K. F. Raffa, and M. G. Turner. 2012. What

Smith, W. K., and G. A. Carter. 1988. Shoot structural effects on

explains landscape patterns of tree mortality caused by bark

needle temperature and photosynthesis in conifers. Am. J. Bot.

beetle outbreaks in Greater Yellowstone? Global Ecol. Biogeog.


21:55667.
Simons, S. B., and T. R. Seastedt. 1999. Decomposition and nitrogen release from foliage of cottonwood (Populus deltoides) and
Russian-olive (Elaeagnus angustifola) in a riparian ecosystem.
Southwest. Nat. 44:25660.
Sims, P. L., and R. T. Coupland.1979. Producers. In R. T. Coupland,
ed., Grassland Ecosystems of the World: Analysis of Grasslands
and Their Uses, pp. 4972. Cambridge Univ. Press, London.
Singer, F. J., L. C. Mark, and R. C. Cates. 1994. Ungulate herbivory of willows on Yellowstone northern winter range. J. Range
Manage. 47:43543.
Singh, J. S., W. K. Lauenroth, and D. G. Milchunas. 1983a. Geography of grassland ecosystems. Prog. Phys. Geogr. 7:4680.

75:496500.
Smith, W. K., and G. N. Geller. 1979. Plant transpiration at high ele
vation: Theory, field measurements, and comparisons of desert
plants. Oecologia 41:10922.
Smith, W. K., and A. K. Knapp. 1990. Ecophysiology of high elevation forests. In C. B. Osmund, L. F. Pitelka, and G. M. Hidy,
eds., Plant Biology of the Basin and Range, pp. 87142. SpringerVerlag, New York.
Smith, W. K., M. J. Germino, T. E. Hancock, and D. M. Johnson.
2003. Another perspective on altitudinal limits of alpine
timberlines. Tree Physiol. 23:110112.
Smith, W. K., M. J. Germino, D. M. Johnson, and K. Reinhardt.
2009. The altitude of alpine treeline: A bellwether of climate
change effects. Bot. Rev. 75:16390.

Singh, J. S., W. K. Lauenroth, R. K. Heitschmidt, and J. L. Dodd.

Smithwick, E. A. H., D. M. Kashian, M. G. Ryan, and M. G. Turner.

1983b. Structural and functional attributes of the vegeta-

2009a. Long-term nitrogen storage and soil nitrogen avail-

tion of northern mixed prairie of North America. Bot. Rev.

ability in post-fire lodgepole pine ecosystems. Ecosystems.

49:11749.

12:792806.

Skinner, Q. D., J. E. Speck, M. A. Smith, and J. C. Adams. 1984.

Smithwick, E. A. H., M. G. Ryan, D. M. Kashian, et al. 2009b. Mod-

Stream water quality as influenced by beaver within grazing

eling the effects of fire and climate change on carbon and ni-

systems in Wyoming. J. Range Manage. 37:14246.

trogen storage in lodgepole pine (Pinus contorta) stands. Global

Skougard, M. G., and J. D. Brotherson. 1979. Vegetational response


to three environmental gradients in the salt playa near Goshen,
Utah County, Utah. Great Basin Nat. 39:4458.
Smith, B. L. 2001. Winter feeding of elk in western North America.
J. Wildlife Manage. 65:17390.

Change Biol. 15:53548.


Sneva, F. A. 1979. The western harvester ant: Their density and hill
size in relation to herbaceous productivity and big sagebrush
cover. J. Range Manage. 32:4647.
Snoke, A. W. 1993. Geological history of Wyoming within the tec-

. 2012. Where Elk Roam: Conservation and Biopolitics of

tonic framework of the North American Cordillera. In A. W.

Our National Elk Herd. Lyons Press, Guilford, Connecticut.

Snoke, Jr., J. R. Steidtmann, and S. M. Roberts, eds., Geology of

Smith, C. C. 1970. The coevolution of pine squirrels (Tamiasciurus)

Wyoming, pp. 256. Memoir 5. Geological Survey of Wyoming,

and conifers. Ecol. Monogr. 40:34971.

Laramie.

. 1975. The coevolution of plants and seed predators. In

Sollins, P., C. C. Grier, F. M. McCorison, et al. 1980. The internal

L.E. Gilbert and P. H. Raven, eds., Coevolution of Animals and

element cycles of an old-growth Douglas-fir stand in western

Plants, pp. 5377. Univ. Texas Press, Austin.


Smith, D. W., R. O. Peterson, and D. B. Houston. 2003. Yellowstone
after wolves. BioScience 53:33040.
Smith, E. L. 1966. Soil-vegetation relationships of some Artemisia
types in North Park, Colorado. Ph.D. diss., Colorado State Univ,
Fort Collins.
Smith, E. W. 1913. Journal of E. Willard Smith 1840. Quart. Oregon
Hist. Soc. 14(3).

Oregon. Ecol. Monogr. 50:26185.


Sonder, L. W. 1959. Soil moisture retention and snow holding capacity as affected by chemical control of big sagebrush (Artemisia tridentata Nutt.). M.S. thesis, Univ. Wyoming, Laramie.
Soul, P. T., and P. A. Knapp. 1996. The influence of vegetation
removal by western harvester ants (Pogonomyrmex owyheei) in
a relict area of sagebrush-steppe in central Oregon. Am. Midl.
Nat. 136:33645.

Smith, F. W., and S. C. Resh. 1999. Age-related changes in produc-

Soul, P. T., P. A. Knapp, and H. D. Grissino-Mayer. 2004. Human

tion and below-ground carbon allocation in Pinus contorta for-

agency, environmental drivers, and western juniper establish-

ests. Forest Sci. 45:33341.

ment during the late Holocene. Ecol. Appl. 14:96112.

Smith, M. A., J. L. Dodd, and J. D. Rogers. 1985. Prescribed burning

Spackman, L. K., and L. C. Munn. 1984. Genesis and morphology

on Wyoming rangeland. Bull. 810. Wyoming Agric. Exp. Sta.,

of soils associated with formation of Laramie Basin (mima-like)

Laramie.
Smith, S. D., and R. S. Nowak. 1990. Ecophysiology of plants in the

mounds in Wyoming. Soil Sci. Soc. Am. J. 48:138492.


Spaeth, K. E. 1981. Successional trends on revegetated Juniperus

intermountain lowlands. In C. B. Osmund, L. F. Pitelka, and

osteosperma communities in north central Wyoming. M.S. the-

G.M. Hidy, eds., Plant Biology of the Basin and Range, pp.179

sis, Univ. Wyoming, Laramie.

241. Springer-Verlag, New York.


Smith, S. D, D. A. Devitt, A. Sala, et al. 1998. Water relations of
riparian plants from warm desert regions. Wetlands 18:68796.

Sperry, J. S., and U. G. Hacke. 2002. Desert shrub water relations


with respect to soil characteristics and plant functional type.
Functional Ecol. 16:36778.

References
Spiering, D. J., and R. L. Knight. 2005. Snag density and use by

Stottlemyer, R., and C. A. Troendle. 1999. Effect of subalpine canopy

cavity-nesting birds in managed stands of the Black Hills

removal on snowpack, soil solution, and nutrient export, Fraser

National Forest. Forest Ecol. Manage. 214:4052.


Spomer, G. G. 1964. Physiological ecology studies of alpine cushion
plants. Physiol. Plant. 17:71724.
Stabler, D. F. 1985. Increasing summer flow in small streams
through management of riparian areas and adjacent vegetation: A synthesis. In R. R. Johnson, C. D. Ziebell, D. R. Patton,
et al., coords., Riparian ecosystems and their management:

Experimental Forest, Colorado. Hydrol. Process. 13:228799.


Stottlemyer, R., C. A. Troendle, and D. Markowitz. 1997. Change in
snowpack, soil water, and streamwater chemistry with elevation during 1990, Fraser Experimental Forest, Colorado. J. Hydrol. 195:11436.
Strahler, A. N. 1969. Physical Geography, 3rd edition. John Wiley
and Sons, New York.

Reconciling conflicting uses, pp. 206l0. First North Ameri-

Strayer, D. L., V. T. Eviner, J. M. Jeschke, and M. L. Pace. 2006.

can Riparian Conference. U.S. Forest Ser. Gen. Tech. Rep.

Under
standing the long-term effects of species invasions.

RM-120.
Stahl, P. D., G. E. Schuman, S. M. Frost, and S. E. Williams. 1998.
Arbuscular mycorrhizae and water stress tolerance of Wyoming
big sagebrush seedlings. Soil Sci. Soc. Am. J. 62:130913.
Stansbury, H. 1852. Exploration and survey of the valley of the
Great Salt Lake of Utah, including a reconnoissance [sic] of a

Trends Ecol. Evol. 21:64551.


Strong, W. E. 1968. A Trip to the Yellowstone National Park in July,
August and September 1875. Univ. Oklahoma Press, Norman.
Sturges, D. L. 1975. Hydrologic relations on undisturbed and converted big sagebrush lands: The status of our knowledge. U.S.
Forest Ser. Res. Paper RM-140.

new route through the Rocky Mountains. Senate special ses-

. 1986. Responses of vegetation and ground cover to spraying

sion, March 1851, pp. 25051. Lippincott, Grambo and Co.,

a high elevation, big sagebrush watershed with 2,4-D. J. Range

Philadelphia.
Stanton, N. L. 1988. The underground in grasslands. Annu. Rev.
Ecol. Syst. 19:57389.
Stapp, P., B. Van Horne, and M. D. Lindquist. 2008. Ecology of

Manage. 39:14146.
. 1993. Soil-water and vegetation dynamics through 20 years
after big sagebrush control. J. Range Manage. 46:16169.
Sturgis, T. 1884. New York Times, December 30, 1884, p. 6.

mammals of the shortgrass steppe. In W. K. Lauenroth and I.C.

Sueltenfuss, J. P., D. J. Cooper, R. L. Knight, and R. M. Waskom.

Burke, eds., Ecology of the Shortgrass Steppe, pp. 13280. Ox-

2013. The creation and maintenance of wetland ecosystems

ford Univ. Press, Oxford.

from irrigation canal and reservoir seepage in a semi-arid land-

Starr, C. R. 1974. Subalpine meadow vegetation in relation to environment at Headquarters Park, Medicine Bow Mountains,
Wyoming. M.S. thesis, Univ. Wyoming, Laramie.
States, J. B. 1968. Growth of ponderosa pine on three geologic
formations in eastern Wyoming. M.S. thesis, Univ. Wyoming,
Laramie.
Steele, R., S. V. Cooper, D. M. Ondov, et al. 1983. Forest habitat
types of eastem Idahowestern Wyoming. U.S. Forest Ser. Gen.
Tech. Rep. INT-144.
Steger, R. E. 1970. Soil moisture and temperature relationships of
six salt desert shrub communities in northcentral Wyoming.
M.S. thesis, Univ. Wyoming, Laramie.

scape. Wetlands 33:799810.


Sun, K. R. 1986. Ranch management of streamside zones. In D. J.
Brosz and D. Rogers, coords., Wyoming Water and Streamside
Zone Conference. Wyoming Water Research Center, Univ. Wyoming, Laramie, and Wyoming Agric. Exp. Sta., Laramie.
Tabler, R. D. 1968. Soil moisture response to spraying big sagebrush
with 2,4-D. J. Range Manage. 21:1215.
Talluto, M., and C. W. Benkman. 2013. Landscape-scale ecoevolutionary dynamics: Selection by seed predators and fire
determine a major reproductive strategy. Ecology 94:130716.
Taylor, D. L., and W. J. Barmore, Jr. 1980. Post-fire succession of
avifauna in coniferous forests of Yellowstone and Grand Teton

Stevens, M. T., M. G. Turner, G. A. Tuscan, et al. 1999. Genetic varia-

National Parks, Wyoming. In R. M. DeGraff, ed., Management

tion in postfire aspen seedlings in Yellowstone National Park.

of Western forests and grasslands for nongame birds, pp. 130

Mol. Ecol. 8:176980.


Stevens, R., and S. B. Monsen. 1988. Hatch winterfat: A quality
shrub for ranges and wildlands. Rangelands 10:1045.
Stinson, K. A. 2005. Effects of snowmelt timing and neighbor density on the altitudinal distribution of Potentilla diversifolia in
western Colorado, USA. Arct. Antarct. Alp. Res. 37:37986.
Stocker, T. F. 2013. The closing door of climate targets. Science
339:28082.
Stohlgren, T. J., K. A. Bull, Y. Otsuki, et al. 1998. Riparian zones as
havens for exotic plant species in the central grasslands. Plant
Ecol. 138:11325.

45. U.S. Forest Ser. Gen. Tech. Rep. INT-86.


Terjung, W. H., R. N. Kickert, G. L. Potter, and S. W. Swarts. 1969.
Energy and moisture balances of an alpine tundra in mid July.
Arct. Alp. Res. 1:24766.
Thatcher, A. P. 1959. Distribution of sagebrush as related to site differences in Albany County, Wyoming. J. Range Manage. 12:5561.
Thilenius, J. F., and D. R. Smith. 1985. Vegetation and soils of an
alpine range in the Absaroka Mountains, Wyoming. U.S. Forest
Ser. Gen. Tech. Rep. RM-121.
Thompson, C. 2008. Water in the Red. In A. Proulx, ed., Red Desert:
History of a Place. Univ. Texas Press, Austin.

Stohlgren, T. J., L. D. Schell, and B. Vanden Heuvel. 1999. How graz-

Thompson, W. W., and F. R. Gartner. 1971. Native forage response

ing and soil quality affect native and exotic plant diversity in

to clearing low-quality ponderosa pine. J. Range Manage. 24:

Rocky Mountain grasslands. Ecol. Appl. 9:4564.

27277.

381

382 References
Thwaites, R. G. 1905. Original Journals of the Lewis and Clark
Exped ition, vol. 7. Dodd, Mead, New York.
Tidwell, W. D. 1975. Common Fossil Plants of Western North America. Brigham Young Univ. Press, Provo, Utah.
Tieszen, L. L., and J. K. Detling. 1983. Productivity of grassland
and tundra. In O. L. Lange, P. S. Nobel, C. B. Osmund, and
H. Ziegler, eds., Physiological Plant Ecology, pp. 173203.
Springer-Verlag, New York.
Tiku, B. L. 1976a. Effect of salinity on the photosynthesis of Salicomia rubra and Distichlis stricta. Physiol. Plant. 37:2328.
. 1976b. Ecophysiological aspects of halophyte zonation in
saline sloughs. Plant Soil 42:35569.

Troendle, C. A., and M. A. Nilles. 1987. The effect of clearcutting


on chemical exports in lateral flow from differing soil depths
on a subalpine forested slope. In R. H. Swanson, P. Y. Bernier,
and P. D. Woodard, eds., Forest Hydrology and Watershed
Management. Intl. Assoc. Hydrol. Sci., Publ. 167, Wallingford,
UK.
Turner, G. T., and D. F. Costello. 1942. Ecological aspects of the
pricklypear problem in eastern Colorado and Wyoming. Ecology 23:41926.
Turner, M. G., W. W. Hargrove, R. H. Gardner, and W. H. Romme.
1994. Effects of fire on landscape heterogeneity in Yellowstone
National Park, Wyoming. J. Veg. Sci. 5:73142.

Tinker, D. B., and D. H. Knight. 2001. Temporal and spatial dy-

Turner, M. G., W. H. Romme, R. H. Gardner, and W. W. Hargrove.

namics of coarse woody debris in harvested and unharvested

1997. Effects of fire size and pattern on early succession in

lodgepole pine forests. Ecol. Model. 141:12549.


. 2004. Snags and coarse woody debris: An important legacy

Yellowstone National Park. Ecol. Monogr. 67:41133.


Turner, M. G., R. H. Gardner and R. V. ONeill. 2001. Landscape

of forests in the Greater Yellowstone Ecosystem. In L. L. Wal-

Ecology in Theory and Practice. Springer-Verlag, New York.

lace, ed., After the Fires: The Ecology of Change in Yellowstone

Turner, M. G., D. B. Tinker, W. H. Romme, et al. 2004. Landscape

National Park, pp. 27998. Yale Univ. Press, New Haven.


Tinker, D. B., W. H. Romme, W. W. Hargrove, et al. 1994. Landscape-scale heterogeneity in lodgepole pine serotiny. Can. J.
Forest Res. 24:897903.
Tolstead, W. L. 1947. Woodlands in northwestern Nebraska. Ecology 28:18088.
Tomback, D. F. 1983. Nutcrackers and pines: Coevolution or co
adaptation? In M. H. Nitecki, ed., Coevolution, pp. 179223.
Univ. Chicago Press, Chicago.
Tomback, D. F., and P. Achuff. 2010. Blister rust and western forest
biodiversity: Ecology, values and outlook for white pines. Forest
Pathol. 40:186225.
Tomback, D. F., A. J. Anderies, K. S. Carsey, et al. 2001. Delayed seed
germination in whitebark pine and regeneration patterns following the Yellowstone fires. Ecology 82:25872600.
Toney, J. L., and R. S. Anderson. 2006. A postglacial palaeoecological record from the Sand Juan Mountains of Colorado USA: Fire,
climate and vegetation history. Holocene 16:50517.
Townsend, J. 1839. Narrative of a journey across the Rocky Moun-

patterns of sapling density, leaf area, and aboveground net primary production in postfire lodgepole pine forests, Yellowstone
National Park (USA). Ecosystems 7:75175.
Turner, M. G., E. A. Smithwick, K. L. Metzger, et al. 2007. In
organic nitrogen availability after severe stand-replacing fire in
the Greater Yellowstone Ecosystem. Proc. Natl. Acad. Sci. USA
104:478289.
Tyrer, S. J., A. L. Hild, B. A. Mealor, and L. C. Munn. 2007. Establishment of native species in soils from Russian knapweed (Acroptilon repens) invasions. Rangeland Ecol. Manage. 60:60412.
Uddin, M. N., D. Caridi, and R. W. Robinson. 2012. Phytotoxic
evaluation of Phragmites australis: An investigation of aqueous
extracts of different organs. Mar. Freshwater Res. 63:77787.
Uhlich, J. W. 1982. Vegetation dynamics of non-grazed and grazed
sagebrush grassland communities in the Big Horn Basin of Wyoming. M.S. thesis, Univ. Wyoming, Laramie.
Unruh, J. D. 1993. The Plains Across: The Overland Emigrants and
the Trans-Mississippi West, 184060. Univ. of Illinois Press,
Urbana.

tains to the Columbia River. Reprinted in R. Thwaites, ed. 1906.

Uresk, D. W., and C. E. Boldt. 1986. Effect of cultural treatment on

Early Western Travels, vol. 21, pp. 107639. Arthur H. Clark,

regeneration of native woodlands on the northern Great Plains.

Cleveland.
Tranquillini, W. 1979. Physiological Ecology of the Alpine Timberline. Springer-Verlag, New York.
Troendle, C. A. 1983. The potential for water yield augmentation
from forest management in the Rocky Mountain region. Water
Resour. Bull. 19:35973.
. 1987. The potential effect of partial cutting and thinning
on streamflow from the subalpine forest. U.S. Forest Ser. Res.
Paper RM-274.
Troendle, C. A., and M. R. Kaufmann. 1987. Influence of forests
on the hydrology of the subalpine forest. In C. A. Troendle,

Prairie Nat. 18:193202.


Uresk, D. W., and W. W. Paintner. 1985. Cattle diets in a ponderosa pine forest in the northern Black Hills. J. Range Manage.
38:44042.
Urness, P. 1989. Why did bison fail west of the Rockies? Utah Sci.
50:17579.
Vale, T. R. 1975. Presettlement vegetation in the sagebrush-grass
area of the intermountain west. J. Range Manage. 28:3236.
. 1978. Tree invasion of Cinnabar Park in Wyoming. Am.
Midl. Nat. 100:27784.
. 2002. The pre-European landscape of the United States: Pris-

M. R. Kaufmann, R. H. Hamre, and R. P. Winokur, technical

tine or humanized? In T. R. Vale, ed., Fire, Native Peoples, and

coords., Management of subalpine forests: Building on fifty

the Natural Landscape, pp. 139. Island Press, Washington, D.C.

years of research, pp. 6878. U.S. Forest Ser. Gen. Tech. Rep.

Van Vuren, D. 1987. Bison west of the Rocky Mountains: A review

RM-149.

of the theories and an explanation. Northwest. Sci. 61:6569.

References
Vander Wall, S. B., and R. P. Balda. 1977. Coadaptation of the

. 1987. Late PleistoceneHolocene environmental changes

Clarks nutcracker and the pinyon pine for efficient seed harvest

in Wyoming: The mammalian record. Illinois State Mus. Sci.

and dispersal. Ecol. Monogr. 47:89111.

Papers 22:33493.

. 1983. Remembrance of seeds stashed. Nat. Hist. (Sept.):


6164.

Walker, M. D., D. A. Walker, T. A. Theodose, and P. J. Webber. 2001.


The vegetation: Hierarchical species-environment relationships.

Vavra, M., R. W. Rice, R. M. Hansen, and P. L. Sims. 1977. Food

In W. D. Bowman and T. R. Seastedt, eds., Structure and Func-

habits of cattle on shortgrass range in northeastern Colorado. J.

tion of an Alpine Ecosystem, pp. 99127. Oxford Univ. Press, New

Range Manage. 30:26163.


Veblen, T. T. 1986a. Treefalls and the coexistence of conifers in subalpine forests of the central Rockies. Ecology 67:64449.
. 1986b. Age and size structure of subalpine forests in the
Colorado Front Range. Bull. Torrey Bot. Club 113:22540.
Veblen, T. T., and D. C. Lorenz. 1986. Anthropogenic disturbance
and recovery patterns in montane forests, Colorado Front
Range. Phys. Geogr. 7:124.
. 1991. The Colorado Front Range: A Century of Ecological
Change. Univ. Utah Press, Salt Lake City.
Veblen, T. T., K. S. Hadley, and M. S. Reid. 1991a. Disturbance and
stand development of a Colorado subalpine forest. J. Biogeogr.
18:70716.
Veblen, T. T., K. S. Hadley, M. S. Reid, and A. J. Rebertus. 1991b.
The response of subalpine forests to spruce beetle outbreaks in
Colorado. Ecology 72:21331.
. 1991c. Methods of detecting past spruce beetle outbreaks in
Rocky Mountain subalpine forests. Can. J. Forest Res. 21:24254.

York.
Wallace, A., and D. L. Nelson. 1990. Wildland shrub die-offs following excessively wet periods: A synthesis. In E. Durant Mc
Arthur, E. M. Romney, S. D. Smith, and P. T. Tueller, compilers,
Symposium on cheatgrass invasion, shrub die-off, and other aspects of shrub biology and management, pp. 8183. U.S. Forest
Ser. Gen. Tech. Rep. INT-276.
Wallace, L. L. 2004. After the Fires: The Ecology of Change in
Yellowstone National Park. Yale Univ. Press, New Haven.
Wallace, L. L., and A. T. Harrison. 1978. Carbohydrate mobilization
and movement in alpine plants. Am. J. Bot. 65:103540.
Walters, J. W., T. E. Hinds, D. W. Johnson, and J. Beatty. 1982.
Effects of partial cutting on diseases, mortality, and regeneration of Rocky Mountain aspen stands. U.S. Forest Ser. Res. Paper
RM-240.
Wardle, P. 1968. Engelmann spruce (Picea engelmannii Engel.)
at its upper limits on the Front Range, Colorado. Ecology
49:48395.

Vlamis, J., A. M. Schultz, and H. H. Biswell. 1964. Nitrogen fixation

Waring, R. H., and G. B. Pitman. 1985. Modifying lodgepole pine

by root nodules of western mountain mahogany. J. Range Man-

stands to change susceptibility to mountain pine beetle attack.

age. 17:7374.

Ecology 66:88997.

Vogt, K. A., C. C. Grier, C. E. Meier, and R. L. Edmonds. 1982.

Washburn, A. L. 1988. Mima mounds, an evaluation of proposed

Mycorrhizal role in net primary production and nutrient cy

origins with special reference to the Puget Lowland. Report of

cling in Abies amabilis ecosystems in western Washington. Ecol-

lnvestigations 29. Washington Div. Geol. Earth Res., Olympia,

ogy 63:37080.
Voight, B. 1982. Lower Gros Ventre Slide, Wyoming, U.S.A. In
B.Voight, ed., Natural Phenomena: Rock Slides and Avalanches,
pp. 11366. Elsevier, Amsterdam.
Vucetich, J., D. W. Smith, and D. R. Stahler. 2005. Influence of
harvest, climate, and wolf predation on Yellowstone elk 1961
2004. Oikos 111:25970.
Wagner, F. H. 2006. Yellowstones Destabilized Ecosystem: Elk
Effects, Science, and Policy Conflict. Oxford Univ. Press, Oxford and New York.
, ed. 2009. Climate Warming in Western North America: Evidence and Environmental Effects. Univ. Utah Press, Salt Lake
City.
Walford, G., G. Jones, W. Fertig, et al. 2001. Riparian and wetland
plant community types of the Shoshone National Forest. U.S.
Forest Ser. Gen. Tech. Rep. RMRS-GTR-85.
Walker, B. L., D. E. Naugle, K. E. Doherty, and T. E. Cornish. 2007.
West Nile Virus and greater sage-grouse: Estimating infection
rate in a wild bird population. Avian Dis. 51:69196.
Walker, D. N. 1982. Early Holocene vertebrate fauna. In G. C. Fri-

Washington.
Watts, J. G., E. W. Huddleston, and J. C. Owens. 1982. Rangeland
entomology. Annu. Rev. Ent. 27:283311.
Waugh, W. J. 1986. Verification, distribution, demography, and
causality of Juniperus osteosperma encroachment at a Big Horn
Basin, Wyoming, site. Ph.D. diss., Univ. Wyoming, Laramie.
Weaver, J. E. 1943. Resurvey of grasses, forbs, and underground
plant parts at the end of the great drought. Ecol. Monogr. 13:
63117.
. 1954. North American Prairie. Johnsen Publishing, Lincoln, Nebraska.
. 1958. Summary and interpretation of underground development in natural grassland communities. Ecol. Monogr. 28:
5578.
. 1965. Native Vegetation of Nebraska. Univ. Nebraska Press,
Lincoln.
. 1968. Prairie Plants and Their Environment. Univ. Nebraska
Press, Lincoln.
Weaver, J. E., and F. E. Clements. 1938. Plant Ecology. McGraw-Hill,
New York.

son and D. J. Stanford, eds., The Agate Basin Site: A Record of

Weaver, T., and D. Klarich. 1977. Allelopathic effects of vola-

the Paleoindian Occupation of the Northwestern High Plains,

tile substances from Artemisia tridentata Nutt. Am. Midl. Nat.

pp.274394. Academic Press, New York.

97:50812.

383

384 References
Weaver, T., and D. Perry. 1978. Relationship of cover type to alti

West, N. E., K. H. Rea, and R. O. Harniss. 1979. Plant demographic

tude, aspect, and substrate in the Bridger Range, Montana.

studies in sagebrush-grass communities of southeastern Idaho.

Northw. Sci. 52:21219.

Ecology 60:37688.

Webb, W. L., S. Szarek, W. Lauenroth, et al. 1978. Primary produc-

Westerling, A. L., H. G. Hidalgo, D. R. Cayan, and T. W. Swetnam.

tivity and water use in native forest, grassland, and desert eco-

2006. Warming and earlier spring increase western U.S. forest

systems. Ecology 59:123947.

wildfire activity. Science 313:94043.

Weisberg, P. J., and W. L. Baker. 1995a. Spatial variation in tree

Westerling, A. L., M. G. Turner, E. A. Smithwick, et al. 2011. Contin-

seedling and krummholz growth in the forest-tundra ecotone

ued warming could transform Greater Yellowstone fire regimes

of Rocky Mountain National Park, Colorado, USA. Arct. Alp.


Res. 27:11629.
. 1995b. Spatial variation in tree regeneration in the foresttundra ecotone, Rocky Mountain National Park, Colorado. Can.
J. Forest Res. 25:132639.
Welch, B. L. 2005. Big sagebrush: A sea fragmented into lakes,
ponds, and puddles. U.S. Forest Ser. Gen. Tech. Rep. 144:1210.
Welch, B. L., F. J. Wagstaff, and J. A. Roberson. 1991. Preference
of wintering sage grouse for big sagebrush. J. Range Manage.
44:46265.
Wells, P. V. 1970. Postglacial vegetational history of the Great
Plains. Science 167:157482.
Welsh, S. L. 1957. An ecological survey of the vegetation of the
Dino
saur National Monument. M.S. thesis, Brigham Young
Univ., Provo, Utah.
Wendtland, K., and J. L. Dodd. 1992. The fire history of Scotts
Bluff National Monument. In D. D. Smith, and J. L. Dodd, eds.,
Recapturing a Vanishing Heritage, pp. 14143. Twelfth North
American Prairie Conference. Univ. Northern Iowa, Cedar
Rapids.
Wentworth, E. N. 1948. Americas Sheep Trails: History and Personalities. Iowa State Univ. Press, Ames.
West, N. E. 1983. Western intermountain sagebrush steppe. In N.
E. West, ed., Ecosystems of the World, vol. 5, Temperate Deserts
and Semi-deserts. Elsevier, New York.
West, N. E. 1988. Intermountain deserts, shrub steppes, and woodlands. In M. B. Barbour and W. D. Billings, eds., North American Terrestrial Vegetation, pp. 20930. Cambridge Univ. Press,
Cambridge.
. 1990. Structure and function of microphytic soil crusts in
wildland ecosystems of arid and semiarid regions. Adv. Ecol.
Res. 20:179223.
. 1991. Nutrient cycling in soils of semiarid and arid regions.

by mid-21st century. Proc. Natl. Acad. Sci. USA 108:1316570.


Weynand, B., W. Behrends, D. Stout, and E. Raper. 1979. Bighorn
river habitat management plan. Internal report. Wyoming
Game and Fish Dept., Cheyenne.
Whicker, A. D., and J. K. Detling. 1998. Ecological consequences of
prairie dog disturbances. BioScience 38:77885.
Whipple, S. A., and R. L. Dix. 1979. Age structure and successional dynamics of a Colorado subalpine forest. Am. Midl. Nat.
101:14248.
Whisenant, S. G. 1990. Changing fire frequencies on Idahos Snake
River plains: Ecological and management implications. In E. D.
McArthur, E. M. Romney, S. D. Smith, and P. T. Tueller, compilers, Symposium on cheatgrass invasion, shrub die-off, and
other aspects of shrub biology and management, pp. 410. U.S.
Forest Ser. Gen. Tech. Rep. INT-276.
White, R. S. 1976. Seasonal patterns of photosynthesis and respiration in Atriplex confertifolia and Ceratoides lanata. Ph.D. diss.,
Utah State Univ., Logan.
Whitlock, C. 1993. Postglacial vegetation and climate of Grand
Teton and southern Yellowstone National Parks. Ecol. Monogr.
63:17398.
Whitman, W. C., and H. C. Hanson. 1939. Vegetation on scoria and
dry buttes in western North Dakota. Ecology 20:45557.
Wick, A. F., P. D. Stahl, L. J. Ingram, et al. 2009. Aggregate-associated
carbon and nitrogen in reclaimed sandy loam soils. Soil Sci.
Soc. Am. J. 73:185260.
Wienk, C. L., C. H. Sieg, and G. R. McPherson. 2004. Evaluating
the role of cutting treatments, fire and seed banks in an experimental framework in ponderosa pine forests of the Black Hills,
South Dakota. Forest Ecol. Manage. 192:37593.
Wiens, J. A., and N. E. McIntyre. 2008. Birds of the shortgrass
steppe. In W. K. Lauenroth and I. C. Burke, eds., Ecology of
the Shortgrass Steppe, pp. 181214. Oxford Univ. Press, Oxford.
Wiens, J. A., G. D. Hayward, H. D. Safford, and C. M. Giffen, eds.

In J. Skujins, ed., Semiarid Lands and Deserts: Soil Resource and

2012. Historical Environmental Variation in Conservation and

Reclamation, pp. 295332. Marcel Dekker, New York.

Natural Resource Management. Wiley-Blackwell, John Wiley

West, N. E., and J. Klemmedson. 1978. Structural distribution of


nitrogen in desert ecosystems. In N. E. West and J. Skujins, eds.,

and Sons, West Sussex, UK.


Wieser, G., F.-K. Holtzmeier, and W. K. Smith. 2014. Ecophysiology,

Nitrogen in Desert Ecosystems, pp. 116. Dowden, Hutchinson,

adaptation, and future survival. In M. Tausz and N. E. Grulke,

and Ross, Stroudsburg, Pennsylvania.

eds., Treeline in a Changing Global Environment. Springer,

West, N. E., and J. Skujins. 1977. The nitrogen cycle in North


American cold-winter semi-desert ecosystems. Oecol. Plant.
12:4553.
West, N. E., and J. A. Young. 2000. Intermountain valleys and lower

New York [in press].


Wight, J. R., and H. G. Fisser. 1968. Juniperus osteosperma in northwest Wyoming: Their distribution and ecology. Wyoming
Agric. Exp. Sta. Sci. Monogr. 7. Univ. Wyoming, Laramie.

mountain slopes. In M. G. Barbour and W. D. Billings, eds.,

Wight, J. R., and E. B. Godfrey. 1985. Predicting yield response

North American Terrestrial Vegetation, 2nd edition, pp. 255

to nitrogen fertilization on Northern Great Plains rangeland.

84. Cambridge Univ. Press, Cambridge.

J.Range Manage. 38:23841.

References
Wight, J. R., and J. T. Nichols. 1966. Effects of harvester ants on production of a saltbush community. J. Range Manage. 19:6871.

Worster, D. 1994. An Unsettled Country: Changing Landscapes of


the American West. Univ. New Mexico Press, Albuquerque.

Willard, B. E., D. J. Cooper, and B. C. Forbes. 2007. Natural regen-

Wotton, B. M., C. A. Nock, and M. D. Flannigan. 2010. Forest fire

eration of alpine tundra vegetation after human trampling: A

occurrence and climate change in Canada. Int. J. Wildlife Fire

42-year data set from Rocky Mountain National Park, Colorado,


USA. Arct. Antarct. Alp. Res. 39:17783.
Williams, J. W., and S. T. Jackson. 2007. Novel climates, no-analog
communities, and ecological surprises. Front. Ecol. Environ.
9:47582.

19:25371.
Wright, H. A., and A. W. Bailey. 1980. Fire ecology and prescribed
burning in the Great Plains: A research review. U.S. Forest Ser.
Gen. Tech. Rep. INT-77.
Wright, H. A., L. Neuenschwander, and C. Britton. 1979. The role

Williams, J. W., S. T. Jackson, and S. E. Kutzbach. 2007. Projected

and use of fire in sagebrush-grass and pinyon-juniper plant

distributions of novel and disappearing climates by 2100 AD.

communities: A state-of-the-art review. U.S. Forest Serv. Gen.

Proc. Natl. Acad. Sci. USA 104:573842.


Wilmers, C. C., and E. Post. 2006. Predicting the influence of wolfprovided carrion on scavenger community dynamics under climate change scenarios. Global Change Biol. 12:4039.
Windell, J. T., B. E. Willard, D. J. Cooper, et al. 1986. An ecological
characterization of Rocky Mountain montane and subalpine
wetlands. U.S. Fish Wildlife Serv. Biol. Rep. 86(11).
Wing, S. L. 1981. A study of paleoecology and paleobotany in the
Willwood Formation (Early Eocene, Wyoming). Ph.D. diss.,
Yale Univ., New Haven.
Wing, S. L., G. J. Harrington, F. A. Smith, et al. 2005. Transient floral change and rapid global warming at the Paleocene-Eocene
boundary. Science 310:99396.
Winward, A. H. 1970. Taxonomy and ecology of big sagebrush.
Ph.D. diss., Univ. Idaho, Moscow.
Wipf, S., V. Stoeckli, and P. Bebi. 2009. Winter climate change in
alpine tundra: Plant responses to changes in snow depth and
snowmelt timing. Climatic Change 94:10521.
Wise, E. K. 2012. Hydroclimatology of the United States intermountain west. Prog. Phys. Geogr. 36:45879.
Wohl, E. 2005. Compromised rivers: Understanding historical
human impacts on rivers in the context of restoration. Ecol.
Soc. http://www.ecologyandsociety.org/vol10/iss2/art2/.
Wolf, E. C., D. J. Cooper, and N. T. Hobbs. 2007. Hydrologic regime and herbivory stabilize an alternative state in Yellowstone
National Park. Ecol. Appl. 17:157287.
Wolfe, C. W. 1973. Effects of fire on a sand hills grassland environment. Tall Timbers Fire Ecol. Conf. 12:24155.
Wollheim, W. M., and J. R. Lovvorn. 1995. Salinity effects on
macro
invertebrate assemblages and water bird food webs in
shallow lakes of the Wyoming High Plains. Hydrobiologia 310:
20723.
Woodhouse, C. A., S. T. Gray, and D. M. Meko. 2006. Updated
streamflow reconstructions for the Upper Colorado River Basin.
Water Resour. Res. 42:WO5415. doi: 10.1029/2005WR004455.
Woodmansee, R. G., J. L. Dodd, R. A. Bowman, et al. 1978. Nitrogen
budget of a shortgrass prairie ecosystem. Oecologia 34:36376.
Woodmansee, R. G., I. Vallis, and J. J. Mott. 1981. Grassland nitrogen. In F. E. Clark and T. Rosswall, eds., Terrestrial Nitrogen
Cycles, pp. 44362. Ecological Bull. 33, Stockholm.

Tech. Rep. INT-58.


Wright, H. E., Jr. 1970. Vegetational history of the Great Plains. In
W. Dort, Jr., and J. K. Jones, Jr., eds., Pleistocene and Recent
Environments of the Central Great Plains, pp. 15772. Univ.
Kansas Press, Lawrence.
Wyeth, N. J. 1899. The Correspondence and Journals of Captain
Nathaniel Jarvis Wyeth 18316. Univ. Press, Eugene, Oregon.
Yellowstone National Park. 1997. Yellowstones Northern Range:
Complexity and Change in a Wildland Ecosystem. National
Park Service, Mammoth Hot Springs, Wyoming.
Yizhaq, H., Y. Ashkenazy, and H. Tsoar. 2009. Sand dune dynamics and climate change: A modeling approach. J. Geophys. Res.
Earth Surface 114: F01023. doi: 10.1029/2008/F001138.
Young, D. L., and J. A. Bailey. 1975. Effect of fire and mechanical
treatment on Cercocarpus montanus and Ribes cereum. J. Range
Manage. 28:49597.
Young, J. A., and F. L. Allen. 1997. Cheatgrass and range science:
19301950. J. Range Manage. 50:53035.
Young, J. A., and R. A. Evans. 1975. Germinability of seed reserves
in a big sagebrush community. Weed Sci. 23:35864.
. 1978. Population dynamics after wildfires in sagebrush
grasslands. J. Range Manage. 31:28389.
. 1981. Demography and fire history of a western juniper
stand. J. Range Manage. 34:5015.
Young, J. A., R. A. Evans, and R. A. Weaver. 1976. Estimating potential downy brome competition after wildfires. J. Range Manage.
29:32225.
Young, J. A., C. D. Clements, and D. Harmon. 2004. Germination
of seeds of Tamarix ramosissima. J. Range Manage. 57:47581.
Young, M. K. 1994. Movement and characteristics of stream-borne
coarse woody debris in adjacent burned and undisturbed watersheds in Wyoming. Can. J. Forest Res. 24:193338.
Young, M. K., D. Haire, and M. A. Bozek. 1994. The effect and extent of railroad tie drives in streams of southeastern Wyoming.
West. J. Appl. Forest. 9:12530.
Youngberg, C. T., and L. Hu. 1972. Root nodules on mountain
mahogany. Forest Sci. 18:21112.
Youngblood, A. P., and W. F. Mueggler. 1981. Aspen community
types on the Bridger-Teton National Forest in western Wyoming. U.S. Forest Ser. Gen. Tech. Rep. INT-272.
Youngblood, A. P., W. G. Padgett, and A. H. Winward. 1985. Ripar

Worrall, J. J., S. B. Marchetti, L. Egeland, et al. 2010. Effects and eti-

ian community type classification of eastern Idahowestern

ology of sudden aspen decline in southwestern Colorado, USA.

Wyoming. R4-Ecol-85-0l. U.S. Forest Ser., Interm. Region,

Forest Ecol. Manage. 260:63848.

Ogden, Utah.

385

386 References
Zachos, J. C., G. R. Dickens, and R. E. Zeebe. 2008. An early Ceno

Zedler, J. B., and S. Kercher. 2004. Causes and consequences of

zoic perspective on greenhouse warming and carbon-cycle dy-

invasive plants in wetlands: Opportunities, opportunists, and

namics. Nature 451:27983.

outcomes. Crit. Rev. Plant Sci. 23:43152.

Zambino, P. J., G. I. McDonald, B. A. Richardson, et al. 2005. Natu-

Zier, J. L., and W. L. Baker. 2006. A century of vegetation change

ral infection of Pedicularis and Castilleja spp. by the white pine

in the San Juan Mountains, Colorado: An analysis using repeat

blister rust fungus Cronartium ribicola in North America. Phytopathology 95:S116.

photography. Forest Ecol. Manage. 228:25162.


Zimmerman, K. M., J. A. Lockwood, and A. V. Latchininsky.

Zamora, B., and P. T. Tueller. 1973. Artemisia arbuscula, A. longiloba

2004. A spatial, Markovian model of rangeland grasshopper

and A. nova habitat types in northern Nevada. Great Basin Nat.

(Orthoptera: Acrididae) population dynamics: Do long-term

33:22541.

benefits justify suppression of infestations? Environ. Entomol.

Zedler, J. B. 2011. Wetlands. In D. Simberloff and M Rejmanek, eds.,


Encyclopedia of Biological Invasions, pp. 698704. Univ. California Press, Berkeley.

33:25766.
Zwingle, E. 1993. Ogallala aquifer: Wellspring of the High Plains.
Natl. Geogr. (March):81109.

Index

Page numbers for entries occurring in figures are followed by an f; those for entries in notes, by an n; and those for entries in tables, by a t.
Bold page numbers indicate pages where a definition can be found.
aboveground net primary productivity (ANPP): in grasslands, 90f; in
lodgepole pine forest, 210f, 340n29
Absaroka Range, 15, 245, 249
Actinobacteria, 64

23739, 237f, 239f; in Greater Yellowstone Ecosystem, 248, 249;


krummholz in, 23133, 232f; major plant communities of, 236;
nitrogen deposition in, 23941, 342nn3637, 343n40; range of
temperatures in, 33; soils of, 232

adaptive radiation, 128

alpine turf, 236, 23839t, 239

adiabatic lapse rate, 33, 173, 327n10, 337n2(ch11)

American beachgrass, 143

Adobe Town, 148f

American elm, 53, 328n8

aeration of soils, 3536

American Indians. See Native Americans

aerenchyma, 6870

American pika, 236, 236f, 242

African grasslands, 89, 94, 97

amphibians: in forests, 179t; in wetlands, 68, 70. See also specific types

Age of Mammals. See Cenozoic Era

anaerobic soils, of wetlands, 6870

aggregating pheromones, 184

Andersen, Chamois, 41

agriculture: area covered by and location of cultivated land, 3, 8f;

Anderson, D. C., 337n2(ch11)

carbon sequestration in, 11; climate change and, 4041; crops

Anderson, E. M., 329n31

of, vs. invasive plants, 329n42; dominant crops in, 325n1(ch1);

Anderson, M. D., 342n20(ch13)

frost-free period and, 33; in grasslands, 83, 92, 93, 100, 1078;

Anderson, S., 347n9

in Laramie Basin, 285, 288, 296, 346n11; limiting factors on, 3;

ANPP. See aboveground net primary productivity

as percentage of state economy, 325n1(ch1); in sagebrush steppe,

antelope. See pronghorn

109; soil characteristics in, 34. See also livestock

antelope bitterbrush, 12021, 143, 156, 158t, 160, 16466, 165f, 176t,

Airport Bench, 287, 288f, 296f

254

Akashi, Yoshiko, 55

ants, harvester, 103, 104f, 139

alder-conifer woodlands, 4749t

apical dominance, 95, 198, 331n7

alfalfa, 325n1(ch1), 346n11

apomixis, 236

alfisols, 174, 322t

aquifers: in Laramie Basin, 286, 298, 346n13; Ogallala, 307, 336n10

alkali flats, 330n20

arctic tundra, 235

alkalinity of soils, 36, 327n17, 335n4

aridisols, 83, 322t

allelopathy, 79

Arizona Monsoon, 31, 196

Allred, B. W., 101, 120

Arkley, R. J., 151

alluvium, in Laramie Basin, 285, 286, 286f, 287

arrowgrass, 72t

alpine pennycress, 235

Ashley, William, 284

alpine sunflower, 236f

aspen forests, woodlands: 8f, 16770, 19698, 196f, 197f; in Black Hills,

alpine tundra, 8f, 23042; adaptations to, 23536, 239; animal

272, 273, 273f; climate change in, 219220; cloning by, 168, 196

species in, 236; avalanches in, 23334, 233f; burrowing animals

97; dominant plant species in, 15859t; 17677t; elk browsing on,

in, 236, 239; climate change in, 24142; dominant plant species

16870, 169f, 256, 257f, 258, 264, 344n60, 344n99; as evergreen

in, 235, 23637, 237f, 23839t; elevation of, 23031; frost in,

tree, 174, 196; fires in, 182, 198, 20910, 261, 262f, 339n99;

387

388 Index
aspen forests (continued)

Berger, J., 343n45

in foothills, 156, 157f, 16770, 167f; genetic diversity in, 197,

Beschta, Robert, 264

339n97; in Greater Yellowstone Ecosystem, 169f, 253, 256, 257f,

Biederman, Joel, 211

258, 26465, 344n60; hydrologic budget of, 205; in Laramie Basin,

big game: current populations of, 24; in foothills and escarpments,

290f, 295; in mountain meadows, 226; sudden aspen decline in,


198, 220, 220f; understory density in, 338n16
aspen atolls, 167, 167f, 226
aspen tent caterpillar, 182
asteroids, 1314

15556, 16871. See also specific types


Big Hollow, 287, 288f, 296f
Bighorn Basin: droughts in, 32; foothills of, 16061, 160f, 162;
frost-free period in, 33; plants in geologic history of, 14, 22, 23f;
sagebrush in, 111; soil crusts in, 126, 126f, 140

atmospheric circulation, 31, 327n5

Bighorn Canyon National Recreation Area, 54

avalanches, 23334, 233f, 253

Bighorn Mountains: alpine tundra in, 230, 231f; average tempera-

badlands, 14748; animal species in, 147; dominant plant species in,

Bighorn River, 52f, 5356, 53f, 56f, 63f

ture changes in, 40; formation of, 15; meadows in, 224f
14748; geologic history of, 17; topography of, 2930, 147

Big Laramie River, 283f, 285, 294f, 296, 347n24

Badlands National Park, 107, 147, 279

big sagebrush. See sagebrush, big

Baker, W. L., 121, 140, 275, 336n39, 342n20(ch13)

Billings, W. D., 225, 342n32, 342n34, 343n45

Bamforth Basin, 287, 288f

bioclimatic envelope, 219

bank storage, 57, 59, 60

bioengineers: beaver as, 57; prairie dogs as, 102

bar(s), in riparian landscapes, 46, 51, 52, 55, 56, 56f

biological control of invasive plants, 65, 79, 329n50


biomass: in alpine tundra, 235, 240; in grasslands, 8485, 86f,
87f, 89, 331n16; in mountain forests, 2012, 202f, 210; in sagebrush steppe, 119. See also net primary productivity
birdfoot sagewort, 135
birds: in alpine tundra, 236; in foothills, 17172; in grasslands, 88t,
333n93; in Greater Yellowstone Ecosystem, 257, 265, 343n45,
344n50; migratory, 68, 330n9; in mountain forests, 178, 179t,
182, 198; roads avoided by, 310, 347n9; seed dispersal by, 17172;
in wetlands, 68, 70. See also specific types
biscuit land. See mima mounds
bison: abundance in grasslands, 334n31; in Black Hills, 27980;
decline in population of, 24, 327n56; effects on grasslands,
9598, 97f, 332n41, 332n50; food preference of, 89; in geologic
history, 13, 325n6(ch2); in Great Plains, 28; hunting of, 24,
332n41; in Laramie Basin, 28485; peak population of, 327n57;
in prairie dog colonies, 97, 97f, 27980, 346n51; in riparian landscapes, 60; sagebrush steppe and, 118, 12627
bitterbrush, antelope, 164, 254
black-backed woodpecker, 182
blackbird, yellow-headed, 67f
blackbody effect of shrubs, 11718
blackfelt snowmold, 225, 226f
black-footed ferret, 10f, 103, 106, 107, 279, 332n76, 346n48
Black Hills, 26681; bark beetle outbreaks in, 271f, 27778; fires in,
187, 27478; foothills of, 166f; formation of, 15, 266; geomorphic
regions of, 267f, 26870, 345n9; grasslands in, 84f, 270, 27374;
invasive plants in, 27879, 279f; management issues in, 27481;
meadows in, 273, 273f; Native Americans in, 266, 345nn12;
plants characteristic of, 26668, 269t, 27074; precipitation in,
267, 270f; topography of, 266, 267f
Black Hills National Forest, 266
black sagebrush. See sagebrush, black
black-tailed prairie dog, 102, 102f, 106, 27879, 280f
BLM. See Bureau of Land Management
Blossey, B., 333n83
bluebunch wheatgrass. See wheatgrass, bluebunch
bluegrass, Kentucky, 79
bluegrass, Sandberg, 29f, 83, 84f, 100, 109f, 110, 12021, 136t, 150f,
164f, 222t, 250, 251f, 289, 291f
bluestem, big, 331n2

bark beetles, 180, 18285, 183; in Black Hills, 27778; climate


change and, 41, 216, 341n63; in Douglas-fir forests, 189; effects
of, 18285, 21011; fires and, 21617, 341n64; fungi in, 182f,
184, 338n38; in Greater Yellowstone Ecosystem, 185, 216,
26263, 340n39; hydrology and nutrient cycling affected by,
211, 340n41; life history of, 18385; in lodgepole pine forests,
185, 216, 338n19; management options for, 21819; as periodic
disturbance, 37; predictors of, 18485, 338n42; species of, 182; in
spruce-fir forests, 195, 211
basin big sagebrush, 111; distribution of, 111, 333n5; greasewood
and, 134; in sand dunes, 143
basins, 5f. See also intermountain basins; specific basins
bastard toadflax, 179
Battaglia, M. A., 346n47
beachgrass, American, 143
bear, grizzly, 26364, 263f; distribution of, 3, 263, 325n2(ch1); pine
seed consumption by, 172, 234, 263; threats to, 26364
Bear Lodge Mountains, 26681; dominant plant species in, 27074;
fires in, 275; foothills of, 166f; formation of, 266; geomorphic
regions of, 270; management issues in, 27481; precipitation in,
267, 270f
Beartooth Mountains: alpine tundra in, 230, 237f, 240f, 342n21;
bryophytes in, 330n29
beaver: impact on riparian landscapes, 5760, 58f, 59f, 329nn2324;
in Laramie Basin, 295; trapping of, 24, 63, 327n54; and willows,
265
Bebb willow, 50f
beetle, Douglas-fir, 182, 189, 189f
beetle, Mediterranean tamarisk, 329n50
beetle, mountain pine, 181f, 18285; in Black Hills, 27778; climate
change and, 216, 341n63; effects of, 18285, 182f, 183f, 219,
340n45; in Greater Yellowstone Ecosystem, 26263
beetle, spruce, 182, 195, 216, 219
beetle, western balsam bark, 182, 195
Belle Fourche River, 266, 280
Belsky, Joy, 95
benches, in Laramie Basin, 286f, 287, 288f, 296f
Benkman, C., 17980
bentonite, 13, 35, 36, 337n28
Berg, A., 151

Index
bluestem, little, 83, 84f, 88, 100, 155, 159t, 163, 166, 268, 27374,
277

caldera, 245
Caldwell, M. M., 115, 331n8

blue spruce, 46, 328n3

cambium, 187f

blue-stain fungus, 182f, 184, 338n38

Camel Rock (in Laramie Basin), 287, 292f

Blumenthal, Dana, 1056, 333n87

Canada thistle, 104

bogs: vs. fens, 77; for geologic history, 14; lack of, in Wyoming,

canarygrass, reed, 79

329n1

canine distemper, 332n76

Bonneville, Lake, 19

canyons, geologic history of, 17. See also specific canyons

Boon, S., 340n41

carbon credits, 11

Bormann, F. Herbert, 200, 31415

carbon dioxide: in Cenozoic Era, 15; current concentrations of,

Boulder Creek, 270


boulder fields, 236, 239
boulders, in alpine tundra, 23637, 239
Bowman, W. D., 342nn3637

328n31; and grasslands, 108; in sagebrush steppe, 119; in soils of


wetlands, 68. See also climate change
carbon flow: in mountain forests, 201, 201f, 210, 211, 340n39,
340n46; in sagebrush steppe, 11819

Boysen Reservoir, 55

carbon refixation, 196

Bradley, B. A., 335n81

carbon sinks, 210, 211, 340n39

braided channels, 46, 51, 51f, 56

Carrera, P. E., 343n44

Breshears, D. D., 337n19

carrying capacity: in mountain meadows, 227; soil characteristics

Brewers sparrow, 310, 347n9

and, 34

Bridger, Jim, 24, 285

Carson, Kit, 28485

Bridger Formation, 326n19

Carter, William A., 106

Bridger Range, 338n4

Cascade Canyon, 19f, 233f

Briske, David, 97

Casper aquifer. See Casper Formation

bristlecone pine, 234, 342n12, 342n15

Casper Formation, 284f, 286, 287, 287f, 293, 293f, 295, 297f

broadleaf cattail, 79

Castle Creek, 273f, 275f

broadleaf woodlands, 4749t

caterpillar, aspen tent, 182

Brown, H. C., 151

cattail, broadleaf, 69f, 70, 74, 79

Brown, J., 260

cattail, hybrid, as invasive species, 79, 331n47

Brown, K., 344n60

cattail, narrowleaf, 69f, 70, 74, 79, 330n47

Brown, M., 340n39

cattle: effects on grasslands, 9596; food preference of, 89; popula-

Brown, P., 275, 276, 346n43

tion of, 25, 327n62. See also livestock

browsing. See grazing and browsing

Cavalry, U.S., 343n41

brucellosis, 344n48

Cenozoic Era, 12f, 1423

bryophytes, 330n29

Centennial Ridge, aspen and Douglas-fir on, 157f

bud sagewort, 135

Centennial Valley, 156, 157f

budworm, western spruce, 182, 189, 189f, 339n65

Central Area of Black Hills, 270, 271f, 345n13

buffalo. See bison

Chain Lakes wetland, 17, 149f

Buffalo Gap National Grassland, 279

Chapman, T., 184

buffaloberry, russet, 175, 274

cheatgrass: in desert shrublands, 138, 140, 141, 336n39; in foothills,

buffaloberry, silver, 46, 48t, 54, 56, 247f

166; in grasslands, 104; introduction of, 123; in Laramie Basin,

buffalo grass, 83, 331nn12

298; in mima mounds, 149; in sagebrush steppe, 12224, 124f,

Bull Lake glacial period, 326n23

334n73

bulrushes, 67f, 70, 71, 72t, 7476, 250, 292

Chimney Rock (in Laramie Basin), 287, 292f

Bultsma, P. M., 332n59

Chisholm, James, 121

bunch grasses, 83

Chokecherry and Sierra Madre Wind Energy Project, 312

Burdick, Laura, 282

chokecherry woodlands, 167. See also woody draws

Bureau of Land Management (BLM): on feral horses, 128; map of

Chugwater Formation, 13, 295, 297f, 345n10

land jurisdiction of, 6f; multiple use on lands of, 3089

Cinnabar Park, 223, 224f, 225, 225f

Burke, I. C., 331n1, 333n94

cinquefoil, shrubby. See shrubby cinquefoil

Burnham, Jennifer, 151

Circle Bar Lake, 148

bur oak, 168, 27273, 272f, 273f

cities, map of, 7f

burrowing animals: in alpine tundra, 236, 239; disturbances caused

Clark, F., 9394

by, 37; in grasslands, 84, 1013; in mima mound formation, 149,

Clark, William, 96, 249

151. See also specific types

Clarks Fork, 307

bushy-tailed woodrat, 22, 22f, 23f

Clarks nutcracker, 17172, 171f, 179, 202


Clean Water Act, 328n2, 329n1

C3 plants: in desert shrublands, 133; in grasslands, 8889

clearcutting: compared to fire, 21115, 340n27; habitat fragmen-

C4 plants: in desert shrublands, 133; in grasslands, 8889

tation by, 311, 311f; of mountain forests, 178. See also timber

cactus, pricklypear, 96, 100, 102

harvesting

389

390 Index
Cleary, M., 119

coyote, 265

Clements, Frederick, 96

crane, sandhill, 328n13

climate, 38; history of, 1223, 32f, 39f, 326n36; vs. weather, 38. See

creeks, map of, 46f. See also specific creeks

also climate change


climate change, 3841; in alpine tundra, 24142; approaches to

Cretaceous-Paleocene Extinction Event, 1314, 325nn89


Cretaceous Period, 12f, 13, 325n3(ch2)

adapting to, 21920, 341n88; approaches to slowing, 40; average

cricket, Mormon, 37, 101, 332n73

temperatures in, 3840, 39f; in Black Hills, 281; computer models

Crist, T. O., 332n71

of, 3840, 39f; current rate of, 15, 3841; in desert shrublands,

cropland. See agriculture

14041; fires and, 41, 26162; in grasslands, 1078; and land

cross-bedding, 287, 287f

management, 31617; in Laramie Basin, 299301, 299f, 300f; in

Crow Creek, 54

mountain forests, 21516, 21920; in mountain meadows, 22829;

crown fires, 37; and bark beetle outbreaks, 217; in Black Hills,

Paleocene-Eocene Thermal Maximum and, 15; and precipitation,


40, 41, 215, 328nn2930; and precipitation-evaporation ratio,

27475, 276, 277f; vs. clearcutting, effects of, 21112; intensity


and severity of, 182, 208, 209f; types of, 217

34; in riparian landscapes, 65; in sagebrush steppe, 120, 121, 124;

cryoturbation, 151; 237, 239, 342n34. See also alpine tundra

tree-ring studies documenting, 23; use of term, 38; in wetlands,

cultivated land. See agriculture

7980

Cunningham, P., 252

climate diagrams, 32f, 289f

curlleaf mountain-mahogany. See mountain-mahogany

cloning: in aspen, 168, 19697; in krummholz, 232; in spruce and

Curtis, J., 327n10

fir, 194, 194f


cloud seeding, 308

cushion plants: in alpine tundra, 235; in desert shrublands, 135,


138f; in foothills, 166

Clovis people, 20, 282

Custer, George A., 266, 273f, 274

Clow, D. W., 340n41

cut banks, 46, 51, 52f

Cluff, G. J., 335n14

cutthroat trout, 26364, 344n85

Clyman, James, 126


coal, formation of, 13, 14, 15, 28788, 326n13. See also energy
extraction
coarse woody debris, 209f, 21115, 213f, 262f, 271f

Dahl, T. E., 330n40


dams: beaver, 5760, 58f, 265; debris, 57, 59f; human-made, 252,
280, 296, 328n16, 329n34

Colby Mammoth Site, 326n30

Daubenmire, R., 185, 338n52, 342n1

cold hardening, 184

Debinski, D., 228

Colorado River: streamflow from Wyoming in, 329n35; water alloca-

debris. See O horizon

tion compacts for, 327n9

debris jams, 57, 59f

Colter, John, 249

DeByle, N., 16970

comandra blister rust, 17879

deciduous woodlands, 16668; in Black Hills, 27273

common reed, 79

decomposition: in grasslands, 92; in mountain forests, 2012, 2078,

Como Bluff, 13
compensatory growth hypothesis, 95

212f; in sagebrush steppe, 119; under snow, 202, 339n3; in


wetlands, 75

Comstock, T., 256, 343n38

decreaser species, 95

cones, serotinous, 17980, 19192, 192f, 339n71

deer, mule: decline in population of, 17071; food preferences of,

conifer woodlands, 4749t, 15859t. See also forests

89; in foothills, 17071; in grasslands, 89; in Greater Yellowstone

conservation banks, 315, 316

Ecosystem, 257; mortality on roads, 310, 347n10; roads avoided

conservation biology, 910, 11

by, 310

conservation finance, 315

Defenders of Wildlife, 264

Conservation Reserve Program, 11, 107, 315

defoliating insects, 182

Continental Divide, 28, 34

denitrification, 207

conveyance losses, 57

desert grassland, 133, 13637t

Cooke, P. G., 54

desert pavement: formation of, 333n10; in mixed desert shrublands,

cool-season plants, 8889

13435; in sagebrush steppe, 112

Cooper, D. J., 265, 330n40

desert playas, 132

Copeland, H., 80

desert shrublands, 8f, 13141, 131f; 134f; adaptations to, 13233;

coppice dunes: in desert shrublands, 13435; in sagebrush steppe,


112, 115f
cottonwood: drought tolerance of, 329n56; elk browsing on, 6061,

animals species in, 139; area covered by, 131; disturbances in,
13940; dominant plant species in, 131, 13637t; ecosystem of,
13739; future challenges in, 14041; major plant communi-

329n33; regeneration after floods, 5253, 5556, 56f, 328n11,

ties of, 13337; precipitation in, 131; resource partitioning in,

328n13

13233; sagebrush steppe compared to, 131, 132, 133, 335n3;

cottonwood, lanceleaf, 53, 328n8


cottonwood, narrowleaf, 46, 51f, 56, 328n10, 328n16, 329n56

soils of, 13134, 139, 335n4; water availability in, 131, 132, 132f,
13739; in Yellowstone National Park, 250

cottonwood, plains, 46, 51f, 52f, 53, 54, 329n56

desert yellowhead, 10f

Coughenour, M., 97

Detling, J., 97

Cox, G., 151

detritus: in grasslands, 90, 92; in mountain forests, 2012

Index
Development by Design, 31516

Yellowstone Ecosystem, 25658, 257f, 259f, 26465; in riparian

Devils Tower National Monument, 268f; establishment of, 3; fires in,

landscapes, 6061; roads avoided by, 310; tree damage by, 168

187f; geologic history of, 266, 345n1; management issues in, 280

70, 169f, 256, 257f, 264, 337n45; winter rangelands and feeding

dikes, 62

of, 17071, 25758; wolves and, 170, 264, 344n99

dinosaurs, 13, 325n6(ch2)

Ellison, L., 96

Dinwoody glacier, 18f

elm, American, 53, 328n8

Diplomystus dentatus, 17f

El Nio, 31

disaggregating pheromones, 184

endangered species: as focus of conservation biology, 910; in Laramie

disturbances, 3638; effects of suppression of, 37, 94; examples of,


3638. See also specific types
Dodge, R. I., 275
Doering, W., 223

Basin, 29798; organizations monitoring, 325n5(ch1); water rights


and, 293, 298, 347n28; of Wyoming, 9. See also specific species
Endangered Species Act: in conservation biology, 9; sage-grouse and,
128; wolves and, 264

doghair forests, 186f, 193, 193f

endemic species, 9

dominance, apical, 95, 198, 331n7

energy extraction: Development by Design and, 316; and ecosystem

Donahue, D., 127

services, 315; through fracking, 3078; and habitat fragmenta-

Dorn, R., 10f, 2425

tion, 31112, 312f, 313f; landscape changes caused by, 3; in

Douglas-fir woodlands and forests, 18890, 188f, 189f; changes

Laramie Basin, 28788; and multiple uses of public lands, 308;

in distribution of, 21920; common plant species in, 17677t;


fires and, 182, 189; in foothills, 16264, 165f; in Grand Teton
National Park, 253; ponderosa pine compared to, 18889;
subspecies of, 338n62; western spruce budworm on, 189f; in
Yellowstone National Park, 189f, 250, 251f

as percentage of state economy, 325n1(ch1); and population


growth, 305; by wind projects, 313f
energy flow, in grasslands, 85, 8992, 91t; in mountain forests,
200202, 201f, 202f, 203f; in sagebrush steppe, 116, 118
Engelmann spruce, 8f, 46, 19396, 194f

Douglas-fir beetle, 182, 189, 189f

Engle, D. M., 332n59

Driese, K. L., 342n3

entisols, 322t

drip irrigation, 62

Eocene Epoch, 14, 15, 245

drought(s): in alpine tundra, 235; bark beetle outbreaks and, 216;

ephemeral leaves, of sagebrush, 116

climate change and, 4041, 65, 328n29; in desert shrublands,

ephemeral streams, 45, 62, 65

13940; grasshoppers during, 100, 12021; in grasslands, 86,

erosion: in badlands, 147; beavers and, 59; in Black Hills and Bear

100, 332n62; history of, 24f, 32, 32f, 327n8; in sagebrush steppe,

Lodge Mountains, 266; in Laramie Basin, 285, 286, 287, 298f;

12021; in sand dunes, 14647; in sudden aspen decline, 219

livestock and, 60, 298f; in mountain forests, after fires, 2089,

drought tolerance: of grassland plants, 86; of invasive plants, 64, 65,


105, 329n56, 333n82

21213, 214, 340n49; in riparian landscapes, 51, 59, 60


escarpments, 15572; big game animals and, 15556, 16871; com-

dunes. See sand dunes

position of, 29; distinctive vegetation on, 29; dominant plant

dusky grouse, 335n96

species on, 156; human impacts on, 172; plant-animal interac-

dust storms, 100

tions on, 16872; woodlands on, 15764

Dutch elm disease, 328n8

EuroAmericans, arrival in Wyoming, 2324, 327n52

dwarf huckleberry, 17375, 190f

eutrophication, 240

dwarf mistletoe, 17879, 178f, 215, 338n19

evaporation: elevation and, 27; as percentage of water consumption,


62; ratio of precipitation to, 34, 35, 203; temperature and, 34. See

ecological regions of Wyoming, 8f; 2830


ecology, 47
ecosystem(s), 5; boundaries of, 7, 313; vs. landscape, use of terms, 7;
major components and interactions of, 56, 9f; map of, 8f

also evapotranspiration
evapotranspiration, 34; in mountain forests, 204; as percentage of
water consumption, 62; potential, 34; temperature in rate of, 34.
See also water availability

ecosystem ecology, 5

even-aged stands, 189, 190

ecosystem management, 31214

evergreen plants: in mountain forests, 174, 175, 196; in sagebrush

ecosystem productivity, net, 211f


ecosystem services, 1011; examples of, 10; in land management,
31415, 314f; value of, 11, 31415. See also specific ecosystem types

steppe, 116
Ewers, B., 119
extinction: climate change and, 40; in Cretaceous-Paleocene

edge forests, habitat fragmentation in, 31112

Extinction Event, 1314, 325n8; habitat fragmentation and, 309;

elevation: of alpine tundra, 230; of Grand Teton National Park, 246,

hunting as cause of, 20

247f, 253, 253f; of Greater Yellowstone Ecosystem, 246, 246f; of


Laramie Basin, 283f, 288; of mountain forests, 173, 174f, 180; of

Fahey, T. J., 339n3

mountain peaks, 27, 173; plant adaptations for, 27, 29f; and pre-

fairy rings, 84, 85f

cipitation, 27, 32f, 327n1; range of, 2728, 4f; and soil character-

Fall, P., 22

istics, 35; and temperature, 27, 33, 173, 327n10; of upper treeline,

farm bills, 329n1; of 2008, 11, 347n17

23031, 342n1, 342n3; Wyoming average, 3

feces deposition, in grasslands, 344n53

elevation map, 4f, 28f

Federal Land Policy and Management Act, 308

elk: in Black Hills, 280; distribution of, 3, 325n2(ch1); food prefer-

fellfields, 236; dominant plant species in, 237, 23839t; formation

ence of, 89; in foothills, 16871; in grasslands, 89; in Greater

of, 239

391

392 Index
fences: around national parks, 280; along roads, 311

21920; beetle outbreaks in (see bark beetle outbreaks); climate

fens, 7779, 77f; vs. bogs, 77; classification of, 78, 330n35; climate

change and, 41; distribution in mountains, 8f, 174f; dominant

change in, 80; dominant plant species in, 68, 7273t, 7779; soils

plants in, 17579, 17677t; edge vs. interior, habitat fragmenta-

of, 6870

tion in, 31112; elevation gradient in, 173, 174f, 180; energy flow

feral horses, in sagebrush steppe, 12728, 127f

in, 200202, 201f, 202f, 203f; expansion into meadows, 223,

Ferrenberg, S. M., 341n63

228, 229; fires in, 18182, 20815, 21618, 216f; future chal-

ferret, black-footed, 10f, 103, 106, 107, 279, 332n76, 346n48

lenges in, 21520; vs. grasslands, 88; historical human impacts

fertilizers, added to grasslands, 92, 93, 106

on, 2526, 21115, 212f, 213f, 30812; hydrology of, 2027, 211,

fescue, Idaho, 50f, 121, 123, 159t, 164, 166, 176t, 194f, 221, 222t,

204f; in Laramie Basin, 29192; major types of, 18598, 338n52;

224f, 250, 252f, 254

management options for, 21719; nutrient cycling in, 17475,

fir, Douglas-. See Douglas-fir

201, 201f, 203f, 2078, 210, 211, 21415; parasitic plants in,

fir, subalpine, 19396, 194f

17879; plant-animal interactions in, 17980, 202; precipitation

fir, white, 339n81

in, 2034, 204f, 21516; slow vs. rapid changes to, 181; soils of,

fire scars, 187, 187f

174, 200202, 2056, 210, 341n7; timber harvesting in, 21115,

fires: bark beetle outbreaks and, 21617, 341n64; in Black Hills, 187,

212f, 213f; tree growth rates in, 200201, 201f, 202f; understory

27478; climate change and, 41, 26162; in desert shrublands,


139, 140; exclusion vs. suppression of, 181, 338n27; in foothills,

density in, 178; in Yellowstone National Park, 189f, 192, 193,


250, 261f. See also specific types of forest

157, 162, 164, 16566, 170; frequency of, 121, 181, 216f; future

forest floor. See O horizon

of, 26162; in grasslands, 37, 85, 98100, 107, 334n32; in Greater

Forest Service, lands administered by, 6f

Yellowstone Ecosystem, 216, 25862, 260f; intensity of, 181, 182,

Fort Union Formation: badlands in, 336n18; coal deposits in, 326n13

208; invasive plants and, 278, 346n47; management options for,

Fossil Butte National Monument, 15, 17f

21718, 276; in mountain forests, 18182, 20815, 21618, 216f;

fossil fuels. See energy extraction

Native American use of, 23, 37, 98, 99; as natural disturbance, 37,

fossil ice wedges, 21f, 151

121, 262; in riparian landscapes, 55; in sagebrush steppe, 37, 110,

Fossil Lake, 15

112, 118, 119, 12123, 122f, 123f, 334n32; in sand dunes, 143;

fossils: in badlands, 147; Cenozoic, 1423; Cretaceous-Paleocene

severity of, 18182, 208, 338n28; vs. timber harvesting, effects

Extinction Event in, 14; Mesozoic, 13; Paleozoic, 1213; tree, 14,

of, 21115, 212f, 213f, 340n27; at upper treeline, 343n45

15, 245, 247f

Firewise recommendations, 218

Fountain Formation, 287, 295

first-order streams, 46

foxtail, Garrison creeping, 29899, 330n44

fish: fossils of, 15, 17f; grazing and, 60, 61f

fracking, 3078

Fish and Wildlife Service, U.S., and endangered species, 325n5(ch1);

fragmentation. See habitat fragmentation

in Laramie Basin, 293; National Wetlands Inventory of, 66; in

Fremont, John, 125, 28485

safe harbor agreements, 316, 347n21; on sage-grouse, 128

Freudenthal, Dave, 306

Fisk, M. C., 342n37

Friggens, M., 41

Fisser, H. G., 337n10

fringed sagewort, 108

flambellate grasshopper, 101f

Frontier Formation, 30f

Flaming Gorge Reservoir, 60, 162, 337n4

Front Range, 241

flood irrigation, 62, 296

frost boils, 237, 239f

floodplains: former, as terraces, 51; glacial outwash, 19, 5152; rapid

frost pockets, 186

changes of, 55; soils of, 45, 51. See also riparian landscapes
floods in riparian landscapes: changes caused by control of, 54, 55,
6162; tree growth after, 5253, 54, 5556, 328n11
food webs: belowground, in grasslands, 34, 90, 91t, 107, 331n16;

fuel breaks, 218


fungi: bark beetles and bluestain, 182f, 184, 338n38; in grasslands,
84, 85f, 89, 331n9; in riparian landscapes, 64; at upper treeline,
234

plant-cow-human vs. plant-human, 9092; in sagebrush steppe,


118; in wetlands, 68. See also energy flow
foothills, 15572; big game animals in, 15556, 16871; dominant

Gage, E., 330n40


Gambel oak, 168

plant species in, 156, 156f, 15859t; evapotranspiration in, 34;

Gannett Peak, 18f

fires in, 157, 162, 164, 16566, 170; formation of, 155; grasslands

Gardner, Dudley, 24

in, 156, 156f, 166; human impacts on, 3, 172; in Laramie Basin,

Gardner saltbush, 132f, 13537, 138f

290f, 29192; loss of habitat in, 168, 17071, 172; mountain-

Garrison creeping foxtail, 29899, 330n44

mahogany shrublands in, 156; plant-animal interactions

gas development. See energy extraction

in, 16872; riparian zones compared to, 155; roads in, 311;

genets, 197

shrublands in, 156, 156f, 157, 16466, 165f, 290f; topography of,

geobotany, 294; of Laramie Basin, 29495, 296f, 297f

15556; woodlands in, 156, 156f, 15764, 16668; of Yellowstone

geologic history, 1226, 12f. See also specific place names

National Park, 250, 251f

Geringer, Jim, 306

Forbis, T. A., 342n21

Geyers willow, 50f

Forelle Limestone, 295, 297f, 298f

geyser basins, of Yellowstone National Park, 251

forests, 173220; adaptations to, 17475, 17980, 181, 182; animal

girdling, 184

species in, 17980, 179t, 19899; area covered by, 8f, 215,

glacial outwash plains, 19, 52, 254, 285

Index
glaciers: and climate change, 40; and floodplains, 19, 5152; forma-

of, 85, 87f, 94, 9697, 332n48; in prairie dog colonies, 9798,

tion of, 19, 20; in Greater Yellowstone Ecosystem, 24648;

1023, 27980; in riparian landscapes, 6061, 61f, 329n30; on

history of, 1723, 326n23; impact on landscapes, 18f, 19, 19f; in

sagebrush, 116, 118, 12528, 333n18, 334n19; in wet meadows,

Laramie Basin, 282, 285, 287; oxygen isotope data from ice cores
of, 328n25; present-day, 20

76
greasewood, 56f, 58f, 74, 74f, 59, 68, 83, 110, 13235; adaptations

gleyed soil horizons, 68

of, 13133; in badlands, 14748, 149f; cheatgrass and, 140;

global warming, use of term, 38. See also climate change

disturbances in, shrublands, 13940; in foothills, 156f, 166; as

gold mining, 63, 266

halophyte, 132; in Laramie Basin, 290f, 292, 294f; seed germina-

Gondwanaland, 13f

tion in, 13334, 335n10; shrublands of, 4749t, 13334, 13637t;

gophers, pocket: bison and, 332n50; in grasslands, 1012; in mima

water availability in, 13739; in Yellowstone National Park, 250

mound formation, 151; in mountain meadows and tundra, 225,

Great Basin wildrye, 164

239, 240f

Great Divide Basin, 28; desert shrublands in, 131f, 133f, 138f; lack of

Goshen Hole, 74

outlet in, 346n1; mud springs in, 14849

Gosiute, Lake, formation of, 1517

greater sage-grouse. See sage-grouse, greater

grama, blue, 36, 8384, 88, 94, 100, 109f, 118, 13435, 136t, 150,

Greater Yellowstone Ecosystem (GYE), 7, 24565; alpine tundra in,

155f, 159t, 268, 273, 289, 291f, 336n21

248, 249; animal species in, 245; bark beetle outbreaks in, 185,

grama, sideoats, 163, 277, 331n2, 332n62

216, 26263, 340n39; boundaries of, 313; dominant plant species

Grand Teton (mountain), 19f, 246, 247f, 251

in, 248; elevation map for, 246f; elk in, 25658, 257f, 26465;

Grand Teton National Park, 25156; avalanches at upper treeline

fires in, 216, 25862; foothills of, 165f, 169f; geologic history of,

of, 233f, 253; boundaries of, 252; dominant plants in, 25254,

24548; management issues in, 25665; mountain meadows in,

253f; elevation of, 246, 247f, 253, 253f; establishment of, 3, 252;

228; riparian zones of, grazing in, 6061, 258; wolves in, 256,

glacial moraines in, 19f; livestock grazing in, 252; mountain


forests in, 19f, 188f
Granite Mountains, 15

26465, 344n93, 344n95. See also specific areas


greenhouse gases: as cause of climate change, 40; in Cenozoic Era,
15. See also carbon dioxide

grasshopper, flambellate, 101f

Green River, 51f

grasshoppers: during droughts, 100, 12021; in grasslands, 100101;

Green River Basin, 15, 32f, 77

number of species of, 332n66; outbreaks of, as disturbances, 37;

Green River Formation, 15, 326n19

in sagebrush steppe, 12021

Green River Lakes, 183f

Grassland National Park, 107, 333n91

green tree retention, 214

grasslands, 8f, 83108, 84f; adaptations to, 8487; agriculture in,

Grimes, K., 327n10

83, 92, 93, 100, 1078; animals of, 8687, 88t; in Black Hills,

Grinnell, George Bird, 343n38

84f, 270, 27374; burrowing animals in, 84, 1013; classification

grizzly bear, 26364, 263f; distribution of, 3, 263, 325n2(ch1); pine

of, 83; climate change and, 1068; conservation and restora-

seed consumption by, 172, 234, 263; threats to, 26364

tion of, 1067, 333n94; desert, 133, 13637t; distribution of, 83;

Gros Ventre landslide, 25556, 255f

dominant plants in, 8384; drought in, 86, 100, 332n62; effects

Gros Ventre Mountains, 25456; dominant plant species in, 25354;

of climate change on, 1078; effects of grazing on, 95f, 105f, 298f

formation of, 15, 25455; landslide in, 25556, 255f

(see also grazing and browsing; livestock grazing); energy flow

Gros Ventre River, 56, 255

(food chain) in, 85, 8992, 91t; fairy rings in, 84, 85f; fertilizers

ground squirrels: in alpine tundra 239; in desert shrublands, 139; in

added to, 92, 93, 106; fires in, 37, 85, 98100, 98f, 107, 334n32;

grasslands, 103; Townsends, 139; Wyoming, 239f

foothill, 156, 156f, 166; future challenges in, 1068; in geologic

groundwater, 3078; percolation to, 206, 3078. See also aquifers

history, 17; grasshoppers in, 100101; invasive plants in, 95,

grouse, species of, 128; dusky, 335n96; sharptailed, 335n96. See also

1036, 333n82; in Laramie Basin, 85f, 99f, 285, 289, 290f, 291f,

sage-grouse, greater

29899; national, 3, 1067; northern mixed prairie, 331n1,

grouse, sage-. See sage-grouse

333n2; nutrient cycling in, 9294, 99; precipitation in, 32, 89;

Gruell, G., 98

resource partitioning in, 8789; sagebrush steppe compared to,

gullies, development of, 24, 5960

117; sagebrush steppe mixed with, 110, 112, 114f; soils of, 34,

GYE. See Greater Yellowstone Ecosystem

83, 8485, 86f, 9293, 108; succession in, 94; tallgrass prairie,

gypsum, 285, 287

331n2; topography of, 84; in Yellowstone National Park, 250,


258, 344n53
Gray, S., 23, 41
Grayson, D. K., 336n38
gray wolf. See wolves
grazing and browsing: in Black Hills, 274, 27980, 345n27; in desert

habitat fragmentation, 30912; causes of, 31012, 313f; climate


change and, 40; effects on wildlife, 30910; in foothills, 172; in
sagebrush steppe, 128
habitat loss: in foothills, 168, 17071, 172; in wetlands, 66, 79,
330n40. See also habitat fragmentation

shrublands, 139, 140, 336nn3536; as disturbance, 327n18;

Hadley, J., 233

effects of, on grasslands, 8889, 9498, 95f, 332n33; and fire,

half-shrubs, 331n3

99100; in foothills, 157, 16162, 16871; and grasshoppers, 101,

halophytes, 36; in desert shrublands, 132; in Laramie Basin, 292,

332n71; in Greater Yellowstone Ecosystem, 6061, 252, 25658;


and invasive plants, 105, 105f; by livestock (see livestock grazing); in mountain meadows, 227, 342n20(ch13); plant tolerance

294f; in playa wetlands, 74; salt excretion by, 132, 132f; on terraces of riparian landscapes, 52
Hansen, A., 265, 314, 343n45, 344n50

393

394 Index
hardpan (caliche), 35

inland saltgrass, 134, 335n14

hardwood draws, 167. See also woody draws

insect(s): climate change and, 41; disturbances caused by, 3738;

Harmony Bench, 287, 288f


Harney Peak, 266

in mountain forests, 180. See also bark beetle outbreaks; specific


types

Hart, R., 25

instream flows, 307, 329n37

harvester ants, 103, 104f, 139

intercalary meristems, 85

Hat Creek Breaks, 163f

interception, 204

Httenschwiler, S., 225

interior forest, fragmentation of, 31112

Hattie, Lake, 288f, 289, 347n24

intermountain basins, 5f; elevation of, 28; evapotranspiration in, 34;

hay: as dominant crop, 325n1(ch1); in Laramie Basin, 285, 288, 296,


29899, 346n11
Hayden, Ferdinand, 25, 148, 149, 249, 251
Hayden Valley, 250, 252f

geologic history of, 13, 15, 16f, 282; map of, 5f; precipitation in,
average annual, 30, 31f; river outlets of, 282, 346n1; topography
of, 4f, 2829. See also specific basins
invasive plants: adaptations of, 104, 105, 333n81; in Black Hills,

Hayman fire (2002), 210, 338n28

27879; control of, 65, 79, 104, 1056, 124, 329n50; vs. crops,

Hedges, Cornelius, 249

329n42; in desert shrublands, 140; drought tolerance of, 64, 65,

herbicides, for invasive plants, 65, 104, 106

105, 329n56, 333n82; fires and, 278, 346n47; in grasslands, 95,

herbivores. See grazing; specific animals

1036, 333n82; in Laramie Basin, 29899; in riparian land-

hibernation, 236, 338n14

scapes, 6365; in sagebrush steppe, 121, 12224; in sand dunes,

Hicke, J., 21617

147, 336n17; in wetlands, 79. See also specific plants

High Savery Reservoir, 308f

inverse texture effect, 35, 142

highways. See roads

irrigation, 6263; area of land with, 62; in Laramie Basin, 80,

Hobbs, T., 265


hogback ridges, 285, 295
Hogback Rim, 26870
Holladay, B., 285

29697, 346n17, 347n24; wetlands supplied by, 67, 74, 76, 80,
346n17. See also agriculture
islands of fertility: in desert shrublands, 139; in sagebrush steppe,
115, 118, 119

Holocene Epoch, 2022, 326n36


Honeycomb Buttes, 147f

Jack Creek, 50f

hopsage, spiny, 112, 134

jackrabbits, 139; white-tailed, 139f

horehound, white, 27879, 279f

Jackson, David, 251

horses, feral: human use of, 23, 24, 127; in sagebrush steppe, 12728,

Jackson, Stephen T., 22, 41

127f; species of, 127


hot springs, in Yellowstone National Park, 251
Houston, D., 258
huckleberry, dwarf, 17375, 190f
humans: arrival in North America, 20, 326n29; early impact on

Jackson, William Henry, 2526, 125f, 249, 251


Jackson Hole, 25156, 254f; dominant plant species in, 25254;
geologic history of, 24546, 248, 248f; riparian landscapes of, 52;
temperature inversions in, 34
Jackson Lake, 252, 343n16

landscapes, 2326; livestock vs. plants in diet of, 9092; popula-

Jackson Lake Dam, 329n34

tion in Wyoming, 305

Jakubos, B., 342n20(ch13)

hummocks, 76

James, J. J., 336n27

Hunt, Wilson Price, 24, 126

Jasper Fire, 277

hunting: extinctions caused by, 20; in Greater Yellowstone

Jenney, W. P., 27475

Ecosystem, 256, 258; by Native Americans, 24, 332n41

Jenny Lake, 19f, 188f

Hutton Lake National Wildlife Refuge, 67f, 74

Jewel Cave National Monument, 270, 277

hybrid cattail, 79, 331n47

Johnson, D., 151

hydraulic fracturing, 3078

Johnson, J. R., 345n9

hydraulic redistribution (lift), 11516, 132

Johnson, K., 2526

hydric soils, 68, 70, 75

Jonah gas field, 316

hydrophytes, 45, 76

juniper, common, 145, 158t, 176t, 183f, 271, 272, 277, 295

hyphae, 202

juniper woodlands, 8f; browsing on, 171f; expansion and infilling

Ice Age. See Pleistocene Epoch

juniper, Rocky Mountain, 23f, 157, 160, 161f, 162, 337n28; browsing

of, 22, 23f, 15762, 15859t, 161f, 337n13


Ice Slough, 7778, 78f

on, 171f

Inceptic Haplocryalf soils, 341n7

juniper, Utah, 22, 23f, 157, 160, 161f, 162

inceptisols, 174, 322t

Jurassic Period, 12f, 13

incised channels, 46, 51f


increaser species, 95

Kashian, D., 209, 340n30

index, site, 339n2

Kay, C., 256

infilling, of juniper, 16162

Kentucky bluegrass, 79

infiltration rates of soils, 3132, 35

kettles, 70, 71f, 248

Ingelfinger, F., 347n9

Keyhole Reservoir, 280

Index
keystone habitats, wetlands as, 66

levees, 62

keystone species, 57; beaver as, 57; prairie dog as, 102

Lewis, Meriwether, 37, 96, 249

Khan, M. A., 335n10

Libby Flats, 228, 232f, 239f

Killpecker Sand Dunes, 142, 143, 144f, 145, 146f

lichens: in alpine tundra, 235; in grasslands, 92, 93f

Kimble, D. S., 344n99

lightning-caused fires, 37; in Black Hills, 275; in grasslands, 98, 98f,

Knopf, F. L., 328n13, 329n33

332n55; in Greater Yellowstone Ecosystem, 259

Krner, C., 241

limber pine. See pine, limber

Kranz, J. J., 338n16

limestone: in Laramie Basin, 285; origin of, 12

Krueger, K., 98

Limestone Plateau, 270, 274

krummholz, 23133, 232f

Lincoln Monument, 292


litter. See O horizon

LAI. See leaf area index

Little Bighorn Battlefield National Monument, 123

Lake Creek, 58f

Little Ice Age, 20, 23, 24f

lake trout, 26364, 344n85

Little Laramie River, 285, 287, 293f, 296

Lamar River valley, 247f, 250, 251f

Little Sage Creek, 50f

lanceleaf cottonwood, 53, 328n8

livestock: arrival of, 25; current number of, 327n62; food preferences

landscape(s): boundaries of, 7; changed vs. unchanged areas of, 3;


vs. ecosystem, use of term, 7; geologic history of, 1226, 12f;
glaciers impact on, 18f, 19, 19f; protective, 314, 347n14; rate of
recent changes to, 3056

of, 89; in human diet, 9092; rise and fall in number of, 25; wolf
predation on, 264
livestock grazing: climate change and, 41; as disturbance, 327n18;
erosion caused by, 60; in grasslands, effects of, 8889, 9498, 95f,

landscape ecology, 7

105f, 298f; historical impacts of, 25, 327n60; in Laramie Basin,

landslides, in Gros Ventre Mountains, 25556, 255f

285, 29596, 298f; on public lands, 127; in riparian landscapes,

Langford, Nathaniel, 343n41

6061, 60f; in sagebrush steppe, 127

La Nia, 31

locust, Rocky Mountain, 101

lapse rate, adiabatic, 33, 173, 327n10, 337n2(ch11)

lodgepole pine. See pine, lodgepole

Laramide orogeny, 13, 14, 15

loess, 19, 142

Laramie (city), 285, 346n13

Logan, J., 263

La Ramie, Jacques, 28283

loosestrife, purple, 79

Laramie Basin, 282301; climate change in, 299301, 299f, 300f; cli-

Lorenz, D., 170

mate of, 28889, 289f; dominant plant species in, 28993; effect

Lower Slide Lake, 255

of livestock grazing in, 298, elevation of, 283f, 288; fires in,

low sagebrush, 112

99f; geobotanical relationships in, 29495, 296f, 297f; geologic

luxury consumption, 210

features of, 282, 284f, 28588, 288f; geologic history of, 282,
285, 287, 346n3; geologic map of, 286f; grasslands in, 85f, 99f,

MacGillivrays warbler, 344n45

285, 289, 290f, 291f, 29899; human history in, 28285, 29596;

Mack, R., 96, 332n41

irrigated land in, 290f; management issues and future challenges

MacMahon, J. A., 337n2(ch11)

in, 295301; maps of, 283f, 286f, 290f; mima mounds in, 149,

Madison Limestone, 12, 270

150f, 151, 293; playa wetlands of, 7475, 74f, 29293; rainshadow

Madson, C., 306

effect in, 288; sagebrush in, 11011, 289, 292; vegetation patterns

mahogany, mountain-. See mountain-mahogany

in, 28993, 290f; wetlands supplied by irrigation in, 80, 29697,

mammals: in alpine tundra, 236; current population of, 24; early

346n17
Laramie Mountains: foothills of, 161f, 164f; forests of, 186f; mima
mounds in, 149
Laramie River, 53, 283f, 285, 296, 297, 328n10. See also Little
Laramie River

human impacts on, 20, 2326; in geologic history, 1317, 20; in


grasslands, 8687, 88t, 89; in Greater Yellowstone Ecosystem,
245; in mountain forests, 17980, 179t, 182, 198; in wetlands,
68. See also specific types
Mammoth Hot Springs, 251

Larocque, Franois A., 327n52

mammoth, 20, 282

Larson, G. E., 345n9

Mann, C., 326n49

Lauenroth, W. K., 90, 96, 331n1, 331n8

Marcott, S. A., 326n36

Laurasia, 13f

Marshall, K., 26465

La Vrendrye brothers, 327n52

marshes, 7074; climate change in, 80; dominant plant species in,

layering, 194, 194f, 232, 342n6

7074, 7273t; invasive plants of, 79; soils of, 6870

leaf area index (LAI), 65, 204, 205f, 210f

Martner, B. E., 327nn78

leafy spurge, 104

mastodons, 282

Lebo, Lake, 15

McIntyre, N. E., 333n93

leks, sage-grouse, 12829

McNaughton, S., 94, 95, 97

Leopold, Aldo, 256, 320, 343n38

Mead, Matt, 306

Leopold, A. Starker, 256

meadows: riparian, 4749t, 54; sagebrush, 22223t; saltgrass, 133,

Leopold report, 25659


Lesica, P., 329n50

134, 13637t. See also mountain meadows; wet meadows


meandering channels, 46, 51, 51f, 56. See also riparian landscapes

395

396 Index
Mears, Brainerd, Jr., 295

distribution of, 111, 112; in foothills, 164; in Grand Teton

Medicine Bow Mountains: alpine tundra and upper treeline in,

National Park, 254, 255f; recovery after disturbances in, 120; in

230, 230f, 231, 232f, 233, 239f, 342n34, 343n45; fens in, 68;
foothills of, 165f; forests in, 190f, 194f, 204, 211, 340n30;

Yellowstone National Park, 251f


mountain-mahogany: curleaf, 157, 160f, 337n4, 337n7, 345n23;

formation of, 15, 326n18; habitat fragmentation by clearcut-

shrublands of, 156, 157, 15859t, 160, 160f; true (alder-leaf),

ting in, 311f; meadows in, 223, 224f, 225, 225f, 227f, 228,

157, 160f, 337n4, 337n7; in Laramie Basin, 289, 291f, 292, 298f,

342n20(ch13)
Mediterranean tamarisk beetle, 329n50

345n23, 346n18
mountain meadows, 22129, 221f; in Black Hills, 273, 273f; climate

Mehringer, P. J., Jr., 334n32

change and, 22829; dominant plant species in, 22123,

meristems, 85

22223t; dry, 22223t; forest expansion into, 223, 228, 229;

Mesozoic Era, 12f, 13, 13f

forests intermingled with, 180; formation of, 223, 225; livestock

Meyer, G., 327n54

grazing on, 227, 341n16, 342n20(ch13); snowglade, 22527; soils

middens, 22, 22f, 23f

of, 223, 225f, 341n7; wet, 75; in Yellowstone National Park, 228,

Middle Rocky Mountains, 30


Mielke, H. W., 332n50

250, 252f
mountain pine beetle, 181f, 18285; in Black Hills, 27778; climate

migratory birds, 68, 330n9

change and, 216, 341n63; effects of, 18285, 182f, 183f, 219,

Mikkelson, K. M., 340n45

340n45; in Greater Yellowstone Ecosystem, 26263. See also bark

Milchunas, D., 90, 96

beetles

Miles, S., 329n50

mountain silver sagebrush. See silver sagebrush

milkvetch, two-grooved, 335n18, 346n22

Muddy Creek, 58f

Millar, C. I., 341n88

mudflats, 7475

mima mounds, 14951, 150f, 293

mud springs, 14849, 149f, 336n19

mining: gold, 63, 266; in Laramie Basin, 285, 287; peat, 79

mule deer: decline in population of, 17071; food preferences of,

Minnelusa Foothills, 270

89; in grasslands, 89; in Greater Yellowstone Ecosystem, 257;

Minshall, G. W., 340n36

mortality on roads, 310, 347n10; roads avoided by, 310Mullen

Miocene Epoch, 17, 325n6(ch2)

Creek, 59f

Missoula, Lake, 19

multiple use on public lands, 3089, 313

Missouri River, 329n35

Munn, L., 151

mistletoe, dwarf, 17879, 178f, 215, 338n19

Munroe, J. S., 341n7

Mitchell, J., 25

mushrooms, in forests, 202, 339n5

Mitton, J. B., 341n63

mustangs, 127. See also horses

mixed desert shrublands, 133, 13435, 135f, 13637t, 139

mustard, Sahara, 336n17

mixed foothill shrublands, 15859t, 16466, 165f

mycoheterotrophs, 175

mixed-grass prairie, 83, 84f; adaptations to, 85; agriculture in, 100;

mycorrhizae: in forests, 174, 175, 175f, 202; mushrooms produced

climate change in, 108; distribution of, 83; dominant plant


species in, 83; drought in, 100; fires in, 99, 99f, 332n59; in

by, 339n5
Myers, W., 295

Laramie Basin, 289, 291f; nutrient cycling in, 92; precipitation


in, 89. See also grasslands

Naftz, D. L., 328n25

Mock, K. E., 339n97

narrowleaf cattail, 79, 330n47

mollisols, 83, 322t

narrowleaf cottonwood, 46, 51f, 56, 328n10, 328n16, 329n56

Monsoon, Arizona, 31, 196

National Academy of Sciences, 258, 308

Mooney, H. A., 342n34

National Climate Assessment program, 40

Moore, J. C., 331n16

National Climate Data Center, 289f, 299

moose, 343n45

National Elk Refuge, 168, 257

Moran, Thomas, 7, 249

National Environmental Policy Act, 309, 315

Morgan, J., 108

National Fire Plan, 218

Mormon cricket, 37, 101, 332n73

national grasslands, 3, 1067

Morrison Formation, 325n6(ch2)

national parks, 6f. See also specific parks

Mortenson Lake National Wildlife Refuge, 293

National Park Service, lands administered by, 6f

mosses, in wetlands, 68, 77, 78, 330n36

National Research Council, 308

Mount Sheridan, 250f

National Weather Service, 40

Mount Washburn, 252f

National Wetlands Inventory, 66

mountain(s), 5f; droughts in, 32; elevation of peaks of, 27, 173;

national wildlife refuges, in Laramie Basin, 29293, 297

evapotranspiration in, 34; fens in, 77; foothills of (see foothills);

Native Americans: arrival of ancestors of, 20; in Black Hills, 266,

formation of, 173; geologic history of, 1317, 173; map of, 5f;

345nn12; fires used by, 23, 37, 98, 99; horses used by, 127; hunt-

marshes in, 70; meadows in (see mountain meadows); other eco-

ing by, 24, 332n41; impact on ecosystems by, 23, 24, 326n49; in

systems affected by, 173; precipitation in, 3032; riparian plants

Laramie Basin, 28285; use of sand dunes, 145; in Yellowstone,

in, 46; topography of, 2830. See also specific mountains


mountain big sagebrush, 111; climate change and, 22829;

249, 343n16
naturalized species, 63, 79

Index
natural regulation policy (Yellowstone National Park), 256,
25758

packrats. See bushy-tailed woodrat


Pahlow strath, 288f

Natural Resources Conservation Service, 128, 204

Painter, E., 95

Nature Conservancy, The, 315

Paintner, W. W., 345n27

Nebraska Sand Hills, 143, 336n10

Paleocene-Eocene Thermal Maximum (PETM), 15

nematodes, 331n9; in grasslands, 86, 89, 98, 331n9

Paleocene Epoch, 13, 1415

net ecosystem productivity (NEP), 211f

paleoecology, 2023, 326n32

net primary productivity (NPP), 201, 327n4; aboveground, 90f,

paleosols, 146

210f, 340n29; in desert shrublands, 335n3; in grasslands, 89, 90f,

Paleozoic Era, 1213, 12f

95f; in forests, 201, 201f, 209, 211f, 339n2, 340nn2931

Palmer Drought Severity Index, 32f

Newton, H., 27475

pantodonts, 14

New West, 306

parasitic fungi, at upper treeline, 234

Niobrara Formation, 29495

parasitic plants, in forests, 17879

nitrogen: in alpine tundra, 23941, 342nn3637, 343n40; in desert

Parkman, Francis, 285

shrublands, 139, 336n28; fixation, 207; in grasslands, 9294, 93f;

parks. See mountain meadows

in forests, 201, 201f, 2078, 207f, 210, 211, 340n36; in riparian

patch-burn grazing system, 100

landscapes, 64; in sagebrush steppe, 119; in sand dunes, 145,

patterned ground. see cryoturbation

336n14; in wetlands, 68

Pawnee National Grassland, 3

nitrogen hotspots, 240

Pearson, J. A., 339n2

nitrogen saturation, 240

peat, in fens, 7778, 330n32

nivation hollows, 114f, 156

peat mining, 79

Niwot Ridge, 342n21, 342n31

Pendall, E., 119

nonriparian wetlands. See wetlands

pennycress, alpine, 235

northern mixed prairie, 331n1, 333n2. See also grasslands

P-E ratio. See precipitation-evaporation ratio

Northern Rocky Mountains, 30

Perry, D., 338n4

northern winter range of Yellowstone National Park, 251f

Persico, L., 327n54

North Fork Fire, 260

Peters, D. P. C., 332n75

North Platte River, 61, 328n13

PETM. See Paleocene-Eocene Thermal Maximum

Northwest Territory sedge, 70, 77f, 78

petrified trees, 245, 247f

Notzold, R., 333n83

Pfeifer, E. M., 340n39

NPP. See net primary productivity

PHACE. See Prairie Heating and Carbon Enhancement Experiment

Nugget Sandstone, 13

pheromones, aggregating and disaggregating, 184

nutcracker, Clarks, 17172, 171f, 179, 202

Phillips, V. D., 342n34

nutrient cycling: in alpine tundra, 235, 23941; in desert shrub-

photographs, early: in Black Hills, 345n15; information on landscape

lands, 139, 336n27; in grasslands, 9294, 99; in forests, 17475,

change in, 2526

201, 201f, 2078, 210, 211, 21415; in sagebrush steppe, 119; in

phragmites (common reed), 79

sand dunes, 14546

phreatophytes, 57, 65

nutrient enrichment, in riparian landscapes, 60, 64

physiological ecology, 5

nutrient resorption, 208, 235

pickleweed, 132f
pika, American, 236, 236f, 242

oak, bur, 168, 27273, 273f

pimple mounds. See mima mounds

oak, Gambel, 168

pine, bristlecone, 234, 342n12, 342n15

oak woodlands, 15859t, 168

pine, limber: arrival of, 22; in foothills, 16264, 164f, 17172; in

obsidian, 249

Laramie Basin, 292, 293f; seed dispersal by birds, 17172; in

Ocean Lake, 74

upper treeline, 233; whitebark pine compared to, 342n3; white

Odum, Eugene, 347n14


offsets, conservation, 316

pine blister rust and, 234; woodlands of, 17677t


pine, lodgepole, 8f, 19093, 190f; adaptations of, 174, 175f; bark

Ogallala aquifer, 307, 336n10

beetle outbreaks and, 182f, 183f, 185, 216, 338n19; distribution

O horizon, 200

of, 8f, 174, 338n4; doghair stands of, 193, 193f; dominant plant

oil development. See energy extraction

species in, 17677t; fires and, 182, 191, 19293, 20810, 209f,

oil shale deposits, formation of, 15

210f, 211f, 213f, 216, 218; forests of, 19093, 190f, 191f; in Grand

old-man-of-the-mountain, 236f

Teton National Park, 253; hydrologic budget of, in forests, 2045,

Old Mans Beard, 272

205f, 206, 340n11; on infertile soils, 190, 250, 339n68; nutrient

Oligocene Epoch, 15

cycling in, in forests, 207; parasitic plants on, 17879; serotinous

olive, Russian, 6365, 63f

cones of, 17980, 19192, 192f, 339n71; timber harvesting in,

open spaces: persistence of, 306, 306f, 307f; value of, 314

213f, 21415; in Yellowstone National Park, 191f, 250. See also

Oregon Trail, 25, 78, 78f, 305, 327n60


Overland Stage Line, 285
Overthrust Belt, 13

forests
pine, pinyon: arrival of, 22; in foothills, 162, 17172, 337n19,
337n22; seed dispersal by birds, 17172

397

398 Index
pine, ponderosa, 8f, 182, 18588, 186f; arrival of, 22, 186; in Black

precipitation-evaporation (P-E) ratio, 34, 35, 203

Hills, 27072, 271f, 274, 275f, 276f, 281; climate change and,

Preisler, H. K., 338n42

281; dominant plant species in forests of, 17677t; Douglas-fir

Preuss, Charles, 125

compared to, 18889; fires and, 182, 186f, 187, 187f, 274, 275f,

pricklypear cactus, 96, 100, 102

276f, 277f, 338n59; in foothills, 155f, 156f, 16264, 163f, 337n26,

primary productivity, net. See net primary productivity

337n28; in Laramie Basin, 292, 293f; in forests, 182, 18588

private lands: area of, 6f; ecosystem management in, 31213, 31516;

pine, whitebark: bark beetle outbreaks and, 26263; limber pine


compared to, 342n3; seed dispersal by birds, 172; in upper

fires on, 218; in Greater Yellowstone Ecosystem, 265; safe harbor


agreements on, 316, 347n21

treeline, 231, 233, 234; white pine blister rust and, 234, 234f,

productivity, net primary. See net primary productivity

26263; in Yellowstone National Park, 250, 250f

pronghorn: in Black Hills, 27980; conservation of, 315f; food

pine beetle, mountain. See mountain pine beetle

preferences of, 89; in geologic history, 13, 17, 325n6(ch2); in

pine cones, serotinous, 17980, 19192, 192f, 339n71. See also pine,

grasslands, 83, 89, 106; mortality on roads, 31011, 347n10; in

lodgepole
Pinedale glacial period, 20, 24648, 326n23
pinedrops, 175
Pine Ridge, 163f

prairie dog colonies, 27980, 346n51; sagebrush consumption


by, 116; wolves and, 265
public lands: area of, 3, 6f, 319; grazing on, 127; multiple use on,
3089

pinyon pine. See pine, pinyon

puffballs, 202

Pioneer Canal, 347n24

pugging, 76

plains cottonwood, 46, 51f, 52f, 53, 54, 329n56

purple loosestrife, 79

Platt, P., 125


Platte River, 22, 297

quaking aspen. See aspen forests, woodlands

playas, desert, 132

Quaternary alluvium, 285

playa wetlands, 7475, 74f; dominant plant species in, 7273t, 7475,

Quaternary Period, 12f, 1721

74f; in Laramie Basin, 7475, 74f, 29293; soils of, 6870


Pleistocene Epoch, 1720, 21f, 325n6(ch2)

radiocarbon dating, 22

Pliocene Epoch, 17

railroads, 24, 25, 285, 305, 329n40

pocket gophers: in alpine and subalpine meadows, 225, 239, 240f;

rainfall. See precipitation

bison and, 332n50; in grasslands, 1012; in mima mound forma-

rainshadow effect, 17, 31, 288

tion, 151; in mountain meadows, 225

ramets, 197

point bars, 46, 51, 52f, 55, 55f, 56f

rangelands. See livestock grazing

pollen record, 20, 21f, 22, 23f, 40, 326n32

Raynolds, W. F., 53

ponderosa pine. See pine, ponderosa

redbeds, 13

ponds, in sand dunes, 145, 146f

Red Buttes, 285

porcupines, 202

Red Desert, 15, 131f, 139f

potholes, 70, 71f, 248

Redman, R. E., 332n59

Powder River Basin: badlands in, 14748, 336n18; coal deposits in,

red (pine) squirrel, 17980, 180f, 192, 234

14, 326n13; escarpments in, 155f; fires in, 98, 98f; geologic his-

red swampfire, 73t, 74, 74f, 132, 134, 137t, 292

tory of, 13; grasshoppers in, 101, 120; habitat fragmentation in,

Red Valley, 270

312; sagebrush steppe in, 111f, 120, 125, 335n81; sage-grouse,

reed, common, 79

130

reed canarygrass, 79

prairie. See grasslands

Rehfeldt, G., 215, 281

prairie dog, black-tailed, 102, 102f, 106, 27879, 280f

Reider, R. G., 151, 223, 326n48

prairie dog, white-tailed, 102

relictual treeline, 241

prairie dogs: in Black Hills, 27879, 279f, 280f; conservation efforts

reptiles: in forests, 179t; in geologic history, 13; in grasslands, 88t

for, 103, 106, 107, 333n78, 333n90; disturbance in grasslands by,


1023, 102f, 332n75; grazing in colonies of, 9798, 97f, 1023,
27980; invasive plants and, 27879, 279f
Prairie Heating and Carbon Enhancement Experiment (PHACE), 108
prairie mounds. See mima mounds
precipitation, 3032; average annual, 30, 31f, 40; in Black Hills and
Bear Lodge Mountains, 267, 270f; climate change and, 40, 41,
215, 328nn2930; cloud seeding and, 308; in desert shrublands,
131; elevation and, 27, 32f, 327n1; in grasslands, 32, 89; infiltration into soil, 3132; in Laramie Basin, 289, 289f, 299301,

reservoirs, 6162; Boysen, 55; development of, 307, 308f; effects of


livestock on, 60; Flaming Gorge, 60, 162, 337n4; functions of, 61;
High Savery, 308f; Keyhole, 280; negative effects of, 61
residential development: in Black Hills, 274, 276; fires in, 218, 276; in
foothills and escarpments, 172, 311; in forests, 218; in Laramie
Basin, 285, 298; in riparian landscapes, 62
resource partitioning: in desert shrublands, 13233; in grasslands,
8789
restoration: of desert shrublands, 141; of grasslands, 1067, 333n94;
of sagebrush steppe, 127

300f; map of, 31f; in forests, 2034, 204f, 21516; in mountain

rest-rotation grazing, 97

meadows, 228; in mountains, 3032; in plant growth, 30, 33;

return flows, 62, 63

range of, 30; ratio of evaporation to, 34, 35, 203; in sagebrush

Reynolds, D. N., 342n21

steppe, 11011, 112, 112f; in sand dunes, 145; seasonality of, 31,

rhizosphere, 70

327n7. See also snow

ribbon forests, 22527, 227f, 231

Index
Richards, J. H., 115, 331n8

Laramie Basin, 11011, 289, 290f, 292; longevity of, 116; old-

Richardson, M. T., 347n24

growth, 118; precipitation in distribution of, 3132, 11011, 112f;

Rickard, W. H., 140

recovery after disturbances in, 12021, 122f, 123f; saline toler-

ridges: hogback, 285, 295; windswept, 156. See also escarpments

ance of, 134, 335n16; varieties of, 11112, 113f; in Yellowstone

riparian landscapes, 4565; area covered by, 45; bank storage in, 57,

National Park, 114f, 250. See also sagebrush islands; sagebrush

59, 60; beavers in, 5760, 58f, 59f; in Black Hills, 272; channels
in, 46, 51, 51f, 5253, 56; climate change in, 65; development of,
46; dominant plant species in, 46, 4749t, 50f, 5257; ecologi-

steppe; other species of sagebrush


sagebrush, black, 110, 112, 114f, 119; distribution of, 112, 156; in
Laramie Basin, 289, 291f, 292, 346n18

cal role of, 45; foothills compared to, 155; future challenges in,

sagebrush, low, 112, 254, 255f

6365; in Greater Yellowstone Ecosystem, 6061, 258; human

sagebrush, mountain big, 111; climate change and, 22829; distribu-

impacts on, 3, 45, 6063, 329n40; invasive animals in, 329n42;

tion of, 111, 112; in foothills, 164; in Grand Teton National Park,

invasive plants in, 6365; landforms in, 4653; in Laramie

254, 255f; recovery after disturbances in, 120; in Yellowstone

Basin, 29293, 290f, 293f, 294f; livestock in, 6061; map of, 46f;

National Park, 251f

vs. nonriparian landscapes, water movement in, 45; percentage

sagebrush, sand, 83, 269t, 333n7

of species dependent on, 45; rapid changes to, 5557; reservoir

sagebrush, silver, 112, 114f, 134f; adaptations of, 116; in grasslands,

construction and, 6162; soils of, 45, 51, 52; succession of plants
in, 54, 56, 56f; water consumption by plants in, 57; wet meadows
in, 75; in Yellowstone National Park, 249, 25051
Ripple, W., 264
Roach, J. L., 333n90
roadless areas, 21718, 309
roads: animal avoidance of, 310, 347n9; animal mortality on,

83; recovery after disturbances in, 120; two varieties of (plains


and mountain), 112
sagebrush, threetip, 112, 115, 156, 157f, 164f, 166, 333n7
sagebrush, Wyoming big, 109f, 110f, 111; distribution of, 111, 112; in
foothills, 164; in grasslands, 83, 290f; recovery after disturbances
in, 120. See also sagebrush, big; sagebrush steppe
sagebrush islands, 11215, 115f

31011; in Black Hills, 274; habitat fragmentation by, 31012,

sagebrush meadows, 22223t

310f; map of, 7f; for timber harvesting, 31112

sagebrush steppe, 8f, 10930; adaptations in, 11517; agriculture

Rockefeller, John D., Jr., 252

in, 109; animals species in, 110, 128; climate change in, 120,

rocks: in alpine tundra, 23637, 239; effects on plant distribution,

121, 124; conservation and restoration of, 127, 12830; desert

29495; as parent material of soils, 3435

shrublands compared to, 131, 132, 133, 335n3; dominant plant

Rocky Mountain juniper, 157, 160, 161f, 162, 171f, 337n28

species in, 11012; drought in, 12021; energy flow in, 116,

Rocky Mountain locust, 101

118; fires in, 37, 110, 112, 118, 119, 12123, 122f, 123f, 334n32;

Rocky Mountain National Park: alpine tundra and upper treeline in,

in Grand Teton National Park, 254; grasshoppers and, 12021;

240, 241, 342n34; fires in, 194

grasslands compared to, 117; grasslands mixed with, 110, 112,

Roman warm period, 21

114f; grazing and, 116, 118, 12528; invasive plants in, 121,

Roosevelt, Theodore, 24

12224; islands of fertility in, 115, 118, 119; loss of, 128; nutrient

root systems: in alpine tundra, 235; in forests, 202, 203f, 206,

availability in, 119; precipitation in, 11011, 112, 112f; recovery

340n19; in grasslands, 8485, 86f, 88; in sagebrush steppe,

after disturbances in, 12024; sage-grouse in, 12830, 129f; soils

11516, 118, 119; in sand dunes, 145

of, 110; topography of, 110, 112; water availability and growth

Rotella, J., 265, 343n45, 344n50


ruminants. See grazing; specific animals
runout zone, 234
rushes, 69f, 72t

in, 11011, 11718; in Yellowstone National Park, 250, 251f


sage-grouse, greater, 12830, 129f; adaptations of, 12829; conservation efforts for, 306; sagebrush consumption by, 116; threats to,
12830

Russell, Osborne, 60, 249

Sage-grouse Initiative, 128

Russian olive, 6365, 63f

sage sparrow, 310, 347n9

rust, comandra blister, 17879

sagewort, birdfoot, 110f, 112f, 131, 131f, 135, 136t, 138f

rust, white pine blister, 17879, 234, 234f, 26263, 342nn1516,

sagewort, bud, 112, 112f, 131f, 133135, 138f

344n78
Ryan, M. G., 340n31

sagewort, fringed, 108, 222t, 224, 250, 278, 280, 289, 331n3
sagewort, threetip. See sagebrush, threetip
Sahara mustard, 336n17

Sacket, D. B., 53

salicornia. See red swampfire

SAD. See sudden aspen decline

salinity of soils, 35, 36, 327n17; vs. alkalinity, 327n17, 335n4; in

safe harbor agreements, 316, 347n21

desert shrublands, 13134, 132f, 139, 335n4; irrigation and,

sage, sand. See sagebrush, sand

6263; in Laramie Basin, 292; in playa wetlands, 7475; in ripar-

sagebrush, basin big, 111, 134f; distribution of, 111, 333n5; in sand
dunes, 143. See also sagebrush, big
sagebrush, big, 10930, 113f, 114f; adaptations of, 11517, 333n18;

ian landscapes, 52
saltbush, Gardner, 13537, 138f
saltbush shrubland, 133, 13537, 13637t, 138f, 139, 336n21

area covered by, 8f, 109, 290f; ecosystem of, 11719; in desert

saltcedar (tamarisk), 54, 56f, 61, 63, 6465, 64f, 329n50

shrublands, 134, 137f; in forests, parasitic plants on, 17879;

salt desert shrubland. See saltbush shrubland

fossil pollen from, 20, 21f; in Grand Teton National Park, 254,

salt flats, 7475, 76, 330n19

255f; grasshopper outbreaks and, 120; in grasslands, 83; grazing

saltgrass, inland, 134, 335n14

on, 125; greasewood and, 134; juniper expansion and, 162; in

saltgrass meadows, 133, 134, 13637t

399

400 Index
salt-secreting glands, 132f

Shinneman, D. J., 140, 275, 336n39, 339n99

salts in soils. See salinity of soils

Shirley Basin: black-footed ferrets in, 333n76; fossil logs in, 14; mima

salt-tolerant plants. See halophytes


sand bars, 52
Sandberg bluegrass. See bluegrass, Sandberg
Sand Creek, 29293
sand dunes, 8f, 14247; adaptations of plants for, 145; animals

mounds in, 149; sagebrush in, 111


shortgrass prairie, 83; agriculture in, 100; distribution of, 83; dominant plant species in, 83, 331n1, 331n8; fires in, 99; nutrient
cycling in, 92, 93. See also grasslands
Shoshone National Forest, 3

species in, 145; disturbances in, 143, 14647; dominant plant

Shoshone River, 54, 5556

species in, 14243, 336n3; factors in location of, 142; formation

shrubby cinquefoil, 48t, 50f, 68, 73t, 75, 75f, 76f, 77f, 167f, 183f,

of, 3435, 142; geologic history of, 13; invasive plants in, 147,
336n17; in Laramie Basin, 287, 293; in mixed desert shrublands,
13435; origin of sand in, 142; ponds in, 145, 146f; precipitation

221f, 223t, 229, 250, 252, 295


shrublands: desert (see desert shrublands); foothill, 156, 156f, 157,
16466; in Laramie Basin, 292, 294f; riparian, 4749t, 5455

in, 145; in sagebrush steppe, 112, 115f; soils of, 14243; stabiliza-

Shuman, B., 3840

tion process in, 143

Sierra Madre: forests in, 196, 196f, 197f; mountain meadows in, 227

sandhill crane, 328n13

Signal Mountain, 71f

sand sage (sagebrush), 83, 269t, 333n7

silver sagebrush. See sagebrush, silver

sandstones: geologic history of, 13; in Laramie Basin, 287, 287f; as


parent material of soils, 3435

silviculture. See timber harvesting


sinks: carbon, 210, 211, 340n39; wetlands as, 79

San Juan Mountains, 219

Sioux, 345n2

saprophytic fungi, 84

site index, 339n2

Saskatoon serviceberry, 16465

Skull Rim, 148f

Satanka Formation, 297f

skunkbush sumac, 47t, 5456, 121, 148, 15556, 158t, 16067, 166f,

Savage Run Wilderness, 309f

176t, 269t, 274, 292f

Schullery, P., 257, 265

slash, 21214, 212f

Schuman, J., 117

Slater, N., 125

scurfpea, slimflower, 145f

slickspot soils, 134

Seastedt, T., 241

slimflower scurfpea, 145f

secondary succession. See succession

Smith, B., 59

second-order streams, 46

Smith, E. W., 284

sedge(s): in alpine tundra, 23839t; in forests, 177t; in grasslands,

Smith, Jedediah, 24

85; in mountain meadows, 222t; in riparian landscapes, 48t; in

Smith, W. K., 225, 231, 233

wetlands, 68, 69f, 7071, 72t, 77, 77f, 78, 250; in Yellowstone

snags, 202f, 214

National Park, 250. See also specific types

Snake River: channelizing of, 6162; dam on, 252, 328n16; vegeta-

sedge, Northwest Territory, 70, 77f, 78


sedge, water, 70, 77f, 78

tion along, 52f, 253; Wild and Scenic Rivers Act protection of
headwaters, 307

Sedgwick, J. A., 328n13, 329n33

Snake River Canyon, 17

sediment deposition: in reservoirs, 61, 307; in riparian landscapes,

snow: in alpine tundra and upper treeline, 23234, 236, 237, 237f,

57
seed dispersal: in foothills, by birds, 17172; in forests, 178, 179; in
riparian landscapes, 54
seed predation: by birds, 171, 179, 202, 234, 262; in forests, 17980,
192; by grizzly bears, 172, 234, 263; at upper treeline, 234
selective timber harvesting. See timber harvesting
selenium: accumulation in plants, 135, 29495, 335n18, 346n22;

241; in avalanches, 23334, 233f; climate change and, 40, 41;


decomposition under, 202, 339n3; in in forests, 173, 2034,
204f, 206; grasslands, 32, 89; infiltration into soil from, 3132;
in mountain meadows, 22527, 226f, 228; in Pleistocene Epoch,
1720; sagebrush distribution affected by, 3132, 11011, 112; in
sagebrush steppe, 32, 111, 112, 114f, 115, 117, 117f; water content
of, 204. See also nivation hollows

in desert shrublands, 135, 139; indicator plants, 135, 29495,

snowglades, 22527, 226f

335n18, 346n22

snowmold, blackfelt, 225, 226f

serotinous cones, 17980, 19192, 192f, 339n71. See also pine,


lodgepole
serviceberry, in Saskatoon and Utah, 156, 158t, 16465, 165f, 176t,
233f, 295, 337n33
Severson, K. E., 338n16
shadscale, 116, 132f; adaptations of, 133; drought and, 13940; in
mixed desert shrublands, 134

Snowy Range, 230f. See also Medicine Bow Mountains


Sodergren Lake, 347n24
sodicity of soils, 327n17, 335n4
sodium in soils. See salinity of soils
soil(s), 3436; alkalinity of, 36, 327n17, 335n4; of alpine tundra and
upper treeline, 232; characteristics of, 3436; climate change and,
108; composition of, 34; of desert shrublands, 13134, 139, 335n4;

shales: geologic history of, 13; mountain meadows and, 224f

development of, 1920, 3435; fires and, 208, 210, 340n28; of

sharptailed grouse, 335n96

forests, 174, 200202, 2056, 210, 341n7; and geologic history,

sheep: in grasslands, food preferences of, 89; population of, 25,

1920; of grasslands, 34, 83, 8485, 86f, 9293, 108; infiltration

327n62
Sheep Mountain, 75f, 293f, 295

rates of, 3132, 35; of mountain meadows, 223, 225f, 341n7;


Ohorizon of, 200; orders of, by vegetation type, 32223t; parent

Index
material of, 3435; in plant growth, 3436, 36f; of riparian

Sully, Alfred, 101

landscapes, 45, 51, 52; of sagebrush steppe, 110; salinity of (see

sunflower, alpine, 236f

salinity); of saltbrush shrubland, 13537, 336n21; of sand dunes,

surface fires, 37; in Black Hills, 274, 275, 276; in Douglas-fir wood-

14243; subgroups of, 32223t, 324t; in temperatures ranges, 33;


transitions between types of, 34; of wetlands, 6870
soil crusts: in desert shrublands, 140, 336n39; in sagebrush steppe,
123, 124, 126, 126f

lands, 18890; intensity and severity of, 182, 208; in ponderosa


pine woodlands, 187. See also fires
sustainable land management. See land management
swampfire, red (salicornia). See red swampfire

solifluction, 23739, 240f, 342n33

Swamp Lake, 66f, 78

Soto, Hernando de, 127

swamps: in geologic history, 14; lack of, in Wyoming, 329n1

Soul, P. T., 337n13

Sweetwater Plateau, 7778

Southern Rocky Mountains, 30

Sweetwater River, 53, 55f, 125f

South Platte River, 328n13

sylvatic plague, 103, 106, 279, 332n76

Spackman, L., 151


sparrow, Brewers, 310, 347n9

tallgrass prairie, 331n2

sparrow, sage, 310, 347n9

Talluto, M., 17980

Spearfish Creek, 270

talus slopes, 236

Spearfish Formation, 13, 270, 345n10

tamarisk. See saltcedar

sphagnum, 14, 77, 250, 330n29, 330n36

Tatman, Lake, 15

spiny hopsage, 112, 13335, 136t, 137, 143, 144f, 335n2, 335n18,

tectonic uplift, 13, 14, 17

336n35, 346n22
springs: in Laramie Basin, 286; in Yellowstone National Park, 251

Telephone Canyon, 292


temperature(s), 3334; average high and low, 33, 33f, 3840,

sprinkler irrigation, 6263

39f, 328n25; elevation and, 27, 33, 173, 327n10; in evapo

spruce, blue, 46, 328n3

transpiration, 34; and fire frequency, 216f; inversions, 3334; in

spruce, Engelmann, 8f, 46, 19396, 194f

Laramie Basin, 288, 289f, 299, 299f, 301; map of, 33f; range of,

spruce, white, 272, 272f, 339n81

33; tolerance of mountain trees, 174; topography and, 3334; at

spruce beetle, 182, 195, 216, 219


spruce-fir forests, 19396, 194f; bark beetle outbreaks in, 195, 211;
dominant plant species in, 17677t; hydrologic budget of, 2045.
See also forests

upper treeline, 233, 342n10


terpenes, 116, 334n19
terraces: in Laramie Basin, 286f; in riparian landscapes, 46, 5152;
along the Snake River, 247f

spurge, leafy, 104

Teton Dam, 252

squirrel, red (pine), 17980, 180f, 192, 234

Teton Range: elevation of, 246, 247f, 248f; formation of, 17, 24546,

squirrels, ground: in alpine, 239f; in desert shrublands, 139; in


grasslands, 103

248f. See also Grand Teton National Park


thinning. See timber harvesting

stands (forest), 338n28; doghair, 193, 193f; even-aged, 189, 190

thistle, Canada, 104

Stansbury, Howard, 285

Thompson, C., 15

Stapp, P., 333n78

Thompson, J., 96, 332n41

state lands, area of, 6f

threetip sagebrush. See sagebrush, threetip

States, J., 163

threetip sagewort. See sagebrush, threetip

state transition hypothesis, 336n35

three-toed woodpecker, 182

steppe, 333n2. See also sagebrush steppe

Thunder Basin National Grassland, 3, 103, 107

Stohlgren, T. J., 341n16

thunderstorms. See lightning

stone nets, 237, 237f

Tiku, B. L., 335n14

stone polygons, 237, 237f, 342n34

timber harvesting: in Black Hills, 274, 278; compared to fire, 21115,

streams: in alpine tundra, 236; ephemeral, 45, 62, 65; map of, 46f;
water movement in, 45. See also riparian landscapes
Strong, W. E., 249

212f, 213f, 340n27; future of, 21516; habitat fragmentation by,


31112, 311f; in multiple use on public lands, 178, 308, 309f;
selective, 21415, 214f, 277f, 341nn5657.

subalpine forests. See spruce-fir forests

toad, Wyoming, 11f, 293, 297

subalpine fir, 8f, 19396, 194f

toadflax, bastard, 179

subalpine meadows. See mountain meadows

topography, 2730, 28f; ecological regions based on, 2830; and

sublimation, 34, 118; from forests, 2047, 205f, 211


succession, primary, in sand dunes, 143, 145, 146
succession, secondary, 94, 325n3(ch1); in aspen, 196-98; in desert

temperature inversions, 3334; and water availability, 34. See also


rainshadow; specific ecosystems and regions
torpor, 175

shrublands, 13940, 336n5; in grasslands, 94, 107; in riparian

tourism, as percentage of state economy, 325n1(ch1)

landscapes, 54, 56, 56f

Townsend, John, 126

succulents, in grasslands, 83, 86

Townsends ground squirrel, 139

suckers, 196. See also aspen forests

transpiration, 204; in forests, 204, 205f, 340n11; in grasslands, 108;

sudden aspen decline (SAD), 198, 219, 220f

in riparian landscapes, 57; in sagebrush, 116, 118, 136. See also

sulfur, in soils of wetlands, 68, 70

evapotranspiration

401

402 Index
Trappers Point, 31011, 347n10
treeline, upper, 23042; adaptations to, 233; advancing of, 24142,

water consumption: evaporation as percentage of, 62; by plants in


riparian landscapes, 57

343n44; avalanches in, 23334, 233f; average temperature at,

water development: cloud seeding, 308; reservoirs, 3078, 308f

233, 342n10; climate change at, 24142; elevation of, 23031,

water-holding capacity of soils, 35

342n1, 342n3; krummholz at, 23133, 232f; soils of, 232; white

water movement, in riparian vs. nonriparian wetlands, 45

pine blister at, 234, 234f. See also alpine tundra

water rights, endangered species and, 293, 298, 347n28

tree-ring studies, 23, 24f, 32

water sedge, 70, 77f, 78

Triassic Period, 12f, 13

Watts, G., 100101

tribal lands, location of, 6f

Waugh, W. J., 162

trona deposits, 15

weather vs. climate, 38

trout, cutthroat, 26364, 344n85

Weaver, John E., vi, 8788, 96

trout, introduced species of, 329n42

Weaver, T., 338n4

trout, lake, 26364, 344n85

weeds. See invasive species of plants

true mountain-mahogany. See mountain-mahogany

West, N. E., 115, 335n3, 336n36

tundra. See alpine tundra

Westerling, A., 26162

Turner, M. G., 340n29

western balsam bark beetle, 182, 195

two-grooved milkvetch, 335n18, 346n22

western spruce budworm, 182, 189, 189f, 339n65

Typic Humicryept soils, 341n7

western wheatgrass, 332n48


West Nile virus, 12930

Uinta Mountains: foothills of, 162; formation of, 15; upper treeline
in, 241
Uncompahgre Plateau, 220f

wetland indicators, 6768, 330n5


Wetland Reserve Program, 80
wetlands, 6680; adaptations to, 6870; animals species in, 68, 70;

Uncompahgre Plateau Partnership, 309

area covered by, 45, 66; biological diversity of, 67; climate change

Union Pacific Railroad, 25, 285, 305

in, 7980; conservation of, 80; construction in, 7677, 80;

uplift, tectonic, 13, 14, 17

dominant plant species in, 6770, 69f, 7273t, 330n5; ecological

Upper Green River Lake, 76f

role of, 45, 66, 329n1; future challenges in, 7980; hydrophytes

Upper Platte River Basin, 299, 300f

used to define, 45; invasive plants in, 79; irrigation water and,

upper treeline. See treeline, upper

67, 74, 76, 80, 346n17; as landscape sinks, 79; of the Laramie

Uresk, D. W., 345n27

Basin, 80, 29697, 346n17; legislation on management of, 328n2;

urine deposition in grasslands, 92, 95, 344n53

loss of, 66, 79, 330n40; seasonal dry periods of, 67, 68, 70; soils

Urness, P., 332n41

of, 6870; types of, 45, 66; water movement in nonriparian vs.

Utah juniper, 22, 23f, 148, 157, 158t, 160, 161f, 162

riparian, 45; in Yellowstone National Park, 250. See also riparian

Utah serviceberry, 16465

landscapes; specific types of wetlands


wet meadows, 7577, 75f, 76f; in alpine tundra, 236, 23839t, 239;

Vale, Thomas, 327n49


Van Vuren, D., 332n41

construction in, 7677; dominant plant species in, 7273t, 7577,


23839t; invasive plants in, 79; soils of, 6870

Veblen, Thomas, 170

wheatgrass, 331n8

vegetation, maps of: in Wyoming, 8f; in Laramie Basin, 290f. See also

wheatgrass, bluebunch, 83, 156, 159t, 164, 165f, 166, 176t, 221, 250,

specific types of vegetation


Vrendrye brothers, 266, 345n2
volcanoes: geologic history of, 13, 15, 17, 20, 245, 325n3(ch2); in
Yellowstone National Park, 20, 245, 326n25
Vore Buffalo Jump, 24

251f
wheatgrass, western, 49t, 83, 84f, 88, 94, 95, 100, 109f, 110, 120,
12123, 13435, 136t, 14748, 155f, 156, 268f, 273, 289, 291f,
294f, 331n8, 332n48, 347n30
whirling disease, 263
whitebark pine: bark beetle outbreaks and, 26263; limber pine com-

Walker Creek, 54f

pared to, 342n3; seed dispersal by birds, 172; at upper treeline,

Wamsutter, Lake, 19, 326n24

231, 233, 234; white pine blister rust and, 234, 234f, 26263; in

warbler, MacGillivrays, 344n45

Yellowstone National Park, 250, 250f

warbler, yellow, 344n50

whitebark pine woodlands, 17677t

warm-season plants, 8889

white fir, 339n81

Wasatch Formation, 326n13

white horehound, 27879, 279f

Washburn, Henry D., 249

white pine blister rust, 17879, 234, 234f, 26263, 342nn1516,

Washburn, Mount, 252f


water allocation: on Colorado River, 327n9; and endangered species,
29798
water availability: climate change and, 4041; contributing factors

344n78
white spruce, 272, 272f, 339n81
white-tailed jackrabbit, 139f
white-tailed prairie dog, 102

in, 30; conveyance losses in, 57; in desert shrublands, 131, 132,

Whitlock, C., 20, 248

132f, 13739; in forests, 2027, 211; in grasslands, 89, 90f; in

Whittlesey, L., 257

mountain meadows, 228; in sagebrush steppe, 11011, 11718;

Wiens, J. A., 333n93

topography in, 34; wind in, 34

Wigand, P. E., 334n32

Index
Wight, J. R., 337n10

Wyeth, Nathaniel, 106, 126

Wild and Scenic Rivers Act, 307, 329n37

Wyoming: area of, 4f; economy of, 31415, 325n1(ch1); ecosystem

wilderness: area of, 3; location of designated, 3, 6f

distribution in, 8f; elevation of, 3, 4f; human population of, 305;

wild horses. See horses, feral

public, private, tribal, and state lands in, 3, 6f; roads and cities

wildlife: in forests, 198, 216; refuges, in Laramie Basin, 29293, 297;


in Yellowstone, 245, 249, 252, 25658. See also specific species
wildrye, Great Basin, 164
wildrye, yellow, 143

of, 7f
Wyoming big sagebrush, 109f, 110f, 111; distribution of, 111, 112; in
foothills, 164; in grasslands, 83; recovery after disturbances in,
120. See also sagebrush, big

Williams, J. W., 41

Wyoming Game and Fish Department, 347n4

willow(s): in alpine tundra, 241; in Greater Yellowstone Ecosystem,

Wyoming Natural Diversity Database, 325n5(ch1), 330n5

25657, 258, 26465; in riparian landscapes, 4546, 50f, 56; in

Wyoming Permanent Mineral Trust Fund, 315

wetlands, 68, 70, 236, 23839t

Wyoming Range, 221f

willow, Bebb, 50f

Wyoming Range Legacy Act, 306

willow, Geyers, 50f

Wyoming toad, 11f, 293, 297

wind: in alpine tundra and upper treeline, 231, 23233, 237, 237f;

Wyoming Wetlands Conservation Strategy, 80

for electrical power, 313f; in foothills, 164; in forest fires, 218;

Wyoming Wildlife and Natural Resources Trust, 306

katabatic, 287, 346n15; in Laramie Basin, 287, 289; in mountain


meadows, 223, 22526; as periodic disturbance, 37; in sagebrush

Xanthoparmelia chlorochroa, 92, 93f

steppe, 112; in sand dune formation, 3435; in water availability,

xerophytes, 8586

34
Wind Cave, 270

yellow cedar, 343n53

Wind Cave National Park: bison in, 97, 97f, 268; black-footed ferrets

yellowhead, desert, 10f

in, 279, 346n48; climate change and other management issues

yellow-headed blackbird, 67f

in, 27481; elk management in, 280; establishment of, 268;

Yellowstone fires of 1988, 25962; aspen forests after, 219, 261, 262f;

fires in, 346n41; intact grasslands in, 97, 107; invasive plants in,

bark beetle outbreaks and, 217, 341n69; and elk, 258; end of, 259;

27879, 279f; prairie dogs in, 97, 97f, 27879, 279f, 280f

extent and severity of, 25960, 260f, 338n28; lessons learned

wind energy projects, and habitat fragmentation, 312, 313f

from, 25961; in lodgepole pine forests, 191f, 192, 209, 210f,

Wind River, 53

211f; net primary productivity after, 209, 210f, 211f, 340n29;

Wind River Basin, 103

nutrient cycling after, 210, 340n36. See also Yellowstone National

Wind River Canyon, 17


Wind River Mountains: bark beetle outbreaks in, 183f; formation of,
15; Gannett Peak, 18f; glaciers in, 20; location, 5f; meadows in,
228

Park
Yellowstone Lake, 248, 344n85
Yellowstone National Park (YNP), 24951; bark beetle outbreaks in,
217, 341n69; dominant plant species in, 25051; early explorers

Wind River Reservation, 329n37

in, 249; ecosystems of, 25051; elk in, 25658, 264; establish-

Wing, Scott, 14

ment of, 3, 249; fens in, 68, 77f; fire history in, 209f, 25859, 261f

winterfat, 132f, 133, 134, 137

(see also Yellowstone fires of 1988); fire management in, 25862;

winter feeding programs, 171, 257

foothills of, 250, 251f; forests in, 189f, 192, 193, 250; fossil trees

wintergreen plants, 174

in, 14, 15, 245, 247f; geologic history of, 245; grasslands in, 250,

winter rangelands: in foothills, 17071; in Greater Yellowstone

258, 344n53; lakes and rivers of, 249; livestock in, 342n20(ch13);

Ecosystem, 251f, 25758

management issues in, 25665, 343n41; mosses vs. vascular

Wisconsin (Pinedale) glacial period, 326n23

plants in, 68; mountain meadows in, 228, 250, 252f; precipita-

wolves, 26465; distribution of, 3, 325n2(ch1); and elk popula-

tion in, average annual, 30; sagebrush in, 114f, 250; temperature

tions, 170, 264, 344n99; in Greater Yellowstone Ecosystem,


256, 26465, 344n93, 344n95; reintroduction of, 170, 245, 258,
26465, 344n93

changes in, average, 40; topography of, 20, 246f, 249; volcanoes
in, 20, 245, 326n25. See also Greater Yellowstone Ecosystem
Yellowstone Plateau, 21920

wood decomposition, 2078, 212f

Yellowstone River, 248, 249, 307

Woodhouse, C. A., 327n9

Yellowstone River valley, 250, 251f

woodpecker, black-backed, 182

yellow warbler, 344n50

woodpecker, three-toed, 182

yellow wildrye, 143

woodrat, bushy-tailed, 22, 22f, 23f

YNP. See Yellowstone National Park

woody draws, 15859t, 166f, 167, 272

Young, J. A., 336n36

Wright, G., 256, 343n38

Young, M. K., 329n40

403

About the Authors

Dennis H. Knight. Professor emeritus at the Univer-

William A. Reiners. Professor emeritus at the Uni-

sity of Wyoming, where he taught ecology and forest

versity of Wyoming, he taught at the University of

management for 35 years. Author of the first edition,

Minnesota and Dartmouth College before moving to

his research focuses on the effects of natural and

Laramie in 1983. His special interests include biogeo-

human-caused disturbances on forest and rangeland

chemistry, ecosystem analysis, and the effects of cli-

ecosystems.

mate change.

George P. Jones. Ecologist and associate director of the

William H. Romme. Professor emeritus, Colorado

Wyoming Natural Diversity Database at the University

State University, he taught at Fort Lewis College for

of Wyoming. His work focuses on the analysis of plant

many years before moving to Fort Collins. Most of his

communities and the provision of ecological informa-

research has been on the effects of fire and other dis-

tion to industry, consulting firms, and local, state, and

turbances on forests, especially in the Greater Yellow-

federal agencies.

stone Ecosystem.

404

Das könnte Ihnen auch gefallen