23.10.2014 Views

December 2012 Number 1 - Utah Native Plant Society

December 2012 Number 1 - Utah Native Plant Society

December 2012 Number 1 - Utah Native Plant Society

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

<strong>December</strong> <strong>2012</strong><br />

<strong>Number</strong> 1<br />

CONTENTS<br />

Proceedings of the Fifth Southwestern<br />

Rare and Endangered<br />

<strong>Plant</strong> Conference<br />

Calochortus nuttallii (Sego lily),<br />

state flower of <strong>Utah</strong>. By Kaye<br />

Thorne.<br />

Calochortiana, a new publication of<br />

the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> . . . . . 3<br />

The Fifth Southwestern Rare and Endangered<br />

<strong>Plant</strong> Conference, Salt Lake<br />

City, <strong>Utah</strong>, March 2009 . . . . . . . . . . 3<br />

Abstracts of presentations and posters<br />

not submitted for the proceedings . . . 4<br />

Southwestern cienegas: Rare habitats<br />

for endangered wetland plants.<br />

Robert Sivinski . . . . . . . . . . . . . . . . . 17<br />

A new look at ranking plant rarity for<br />

conservation purposes, with an emphasis<br />

on the flora of the American<br />

Southwest. John R. Spence . . . . . . . 25<br />

The contribution of Cedar Breaks National<br />

Monument to the conservation<br />

of vascular plant diversity in <strong>Utah</strong>.<br />

Walter Fertig and Douglas N. Reynolds<br />

. . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

Studying the seed bank dynamics of<br />

rare plants. Susan Meyer . . . . . . . . . 46<br />

East meets west: Rare desert Alliums<br />

in Arizona. John L. Anderson . . . . . . 56<br />

Spatial patterns of endemic plant species<br />

of the Colorado Plateau. Crystal<br />

Krause . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

Continued on page 2<br />

Copyright <strong>2012</strong> <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. All Rights Reserved.


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>, PO Box 520041, Salt Lake<br />

City, <strong>Utah</strong>, 84152-0041. www.unps.org<br />

Editor: Walter Fertig (walt@kanab.net),<br />

Editorial Committee: Walter Fertig, Mindy Wheeler,<br />

Leila Shultz, and Susan Meyer<br />

Copyright <strong>2012</strong> <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. All Rights<br />

Reserved. Calochortiana is a publication of the <strong>Utah</strong><br />

<strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>, a 501(c)(3) not-for-profit organization<br />

dedicated to conserving and promoting stewardship<br />

of our native plants.<br />

CONTENTS, continued<br />

Biogeography of rare plants of the Ash Meadows National Wildlife Refuge, Nevada. Leanna Ballard . . . . . . . . . 77<br />

Assessing vulnerability to climate change among the rarest plants of Nevada’s Great Basin. Steve Caicco . . . . . 91<br />

Sentry milkvetch (Astragalus cremnophylax var. cremnophylax) update. Janice Busco . . . . . . . . . . . . . . . . . . . . 106<br />

A tale of two single mountain alpine endemics: Packera franciscana and Erigeron mancus. James F. Fowler,<br />

Carolyn Hull Sieg, Brian M. Cassavant, and Addie E. Hite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110<br />

Long-term population demographics and plant community interactions of Penstemon harringtonii, an endemic<br />

species of Colorado’s western slope. Thomas A. Grant III, Michele E. DePrenger-Levin, and Carol Dawson . . 115<br />

Conservation and restoration research at The Arboretum at Flagstaff. Kristin E. Haskins and Sheila Murray . . . . 120<br />

The digital Atlas of <strong>Utah</strong> <strong>Plant</strong>s: Determining patterns of biodiversity and rarity. Leila M. Shultz, R. Douglas<br />

Ramsey, Wanda Lindquist, and C. Garrard. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

Molecular genetic diversity and differentiation in Clay phacelia (Phacelia argillacea Atwood: Hydrophyllaceae).<br />

Steven Harrison, Susan E. Meyer, and Mikel Stevens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127<br />

A taxonomic revision of Astragalus lentiginosus var. maricopae and Astragalus lentiginosus var. ursinus two<br />

taxa endemic to the southwestern United States. Jason Alexander . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

Ecology of Rusby’s milkvetch (Astragalus rusbyi), a rare endemic of northern Arizona ponderosa pine forests.<br />

Judith D. Springer, Michael T. Stoddard, Daniel C. Laughlin, Debra L. Crisp, and Barbara G. Phillips . . . . . . 157<br />

Long-term responses of Penstemon clutei (Sunset Crater beardtongue) to root trenching and prescribed fire:<br />

Clues for population persistence. Judith D. Springer, Peter Z. Fulé, and David W. Huffman . . . . . . . . . . . . . . . . 164<br />

¡Viva thamnophila! Ecology of Zapata bladderpod (Physaria thamnophila), an Endangered plant of the Texas-<br />

Mexico borderlands. Dana M. Price, Christopher F. Best, Norma L. Fowler, and Alice L. Hempel . . . . . . . . . 172<br />

Intraspecific cytotype variation and conservation: An example from Phlox (Polemoniaceae). Shannon D.<br />

Fehlberg and Carolyn J. Ferguson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189<br />

Prioritizing plant species for conservation in <strong>Utah</strong>: Developing the UNPS rare plant list. Walter Fertig . . . . . . . . 196<br />

2


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Calochortiana, a New Publication of the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Hundreds of scientific journals already exist for the dissemination of research on botany and ecology (including<br />

several fine publications based in <strong>Utah</strong> and the west). Nonetheless, space and financial constraints prevent many useful<br />

papers from being published in first and second-tier journals, relegating such work to the gray literature. In June<br />

<strong>2012</strong>, the board of the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> (UNPS) recognized the need for a peer-reviewed, electronic journal<br />

for unpublished gray-literature reports that pertain to <strong>Utah</strong> botany and vegetation. The board voted to establish an<br />

annual, technical journal that would complement its bimonthly member’s magazine, the Sego Lily. The objective of<br />

the new publication, named Calochortiana (“of or relating to Calochortus or Sego Lily”, the state floral emblem of<br />

<strong>Utah</strong>), is to provide a forum for professional and amateur scientists to share their findings on <strong>Utah</strong> botany and ecology<br />

with their colleagues. Calochortiana will focus primarily on monitoring or status surveys of rare species, seed<br />

propagation protocols, floristic checklists, genetic studies, vegetation mapping, natural history research, or other topics<br />

that might not otherwise be accepted in existing journals. All submissions will be peer-reviewed and the journal<br />

made available for free on the UNPS website. The journal is put together by an all-volunteer editorial board, though<br />

supported by UNPS. Readers, of course, are encouraged to show their appreciation by becoming members of UNPS!<br />

This first issue of Calochortiana contains papers presented at the 5th Southwestern Rare and Endangered <strong>Plant</strong><br />

Conference, hosted by UNPS in March 2009. These papers were originally intended for publication by the US Forest<br />

Service as part of a proceedings volume. Unfortunately, staff changes, budget shortfalls, and new policy review requirements<br />

greatly delayed publication of the proceedings by the Forest Service. In October <strong>2012</strong>, UNPS assumed<br />

responsibility for disseminating the conference papers to help launch its new journal. The second issue of Calochortiana<br />

will be published on the UNPS website (www.unps.org) in the fall of 2013. Submissions for that issue will be<br />

accepted through 30 April 2013. For more information, please contact me (walt@kanab.net). - Walter Fertig<br />

The Fifth Southwestern Rare and Endangered <strong>Plant</strong> Conference<br />

Salt Lake City, <strong>Utah</strong>, March 2009<br />

In late 2007, botanists in the southwestern United States began discussions about holding a region-wide rare plant<br />

conference modeled after the Fourth Southwestern Rare and Endangered <strong>Plant</strong>s meeting held in Las Cruces, New<br />

Mexico in 2004. It was widely acknowledged through the botanical grapevine that it ought to be <strong>Utah</strong>’s turn to host<br />

the event. Mindy Wheeler, who was chair of the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> (UNPS) at the time, proposed that the<br />

<strong>Society</strong> take the lead in organizing a conference, slated for early spring 2009. UNPS already had experience with cohosting<br />

the annual state rare plant meeting with Red Butte Garden, so how hard could a regional conference be?<br />

Without going into the gory details, the months of developing an agenda, finding a venue, creating a website, signing<br />

up sponsors, sending out invitations to speakers and attendees, organizing field trips, hiring caterers, and completing<br />

hundreds of other tasks all just seemed to whisk by. On the evening of March 16, 2009, UNPS was proud to host<br />

the first event of the Fifth Southwestern Rare and Endangered <strong>Plant</strong> Conference - an informal mixer at historic Fort<br />

Douglas on the campus of the University of <strong>Utah</strong> in Salt Lake City. Fortified by good food, fine spirits, and excellent<br />

company, the organizers and participants of the conference were off to a good start.<br />

The conference officially began the following morning. Noel Holmgren, curator emeritus of the New York Botanical<br />

Garden, gave the keynote address in which he briefly outlined the history of the Garden’s Intermountain Flora<br />

project and described patterns of species richness and endemism in the Great Basin, Colorado Plateau, and the rest of<br />

the Southwest. UNPS presented Noel and Pat Holmgren with hand-crafted lanyards (for their hand lenses) in appreciation<br />

of their decades of work on the Intermountain Flora.<br />

Over the next three days, 36 additional speakers gave presentations or workshops and an additional 20 posters<br />

were displayed at an evening reception. Presentations covered a variety of topics, ranging from seedling ecology and<br />

rare plant biology to distributional modeling, impacts of climate change, plant biogeography, and fire ecology.<br />

The conference concluded with a Friday field trip to Stansbury Island along the south side of the Great Salt Lake.<br />

Despite the unusually warm temperatures of mid-March, relatively few plants were flowering, though attendees were<br />

treated to a display of violet buttercup (Ranunculus andersonii var. andersonii) in bloom.<br />

3


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

All told, over 150 botanists attended the week-long conference. Much of the success of the conference could be<br />

attributed to the hard work of the planning and program committees, both chaired by Mindy Wheeler with the able<br />

assistance of Bill Gray, Ann Kelsey, Bill King, Therese and Larry Meyer, Robert and Susan Fitts, Loreen Allphin,<br />

Rita (Dodge) Reisor, and Leila Shultz. A number of volunteers from UNPS and Red Butte Garden helped with registration,<br />

food, and behind the scenes work, including Elise Erler, Tony Frates, Celeste Kennard, Kipp Lee, Bill Nelsen,<br />

Kody Wallace, Sue Budden, Pamela and Robert Hilbert, Allene Keller, Jena Lewinsohn, Marilyn Mead, and Bev<br />

Sudbury. Artist Laura Call Gastinger provided a beautiful painting of Dwarf bearclaw poppy (Arctomecon humilis)<br />

for the conference program and souvenir mug. The following corporate and institutional sponsors assisted financially<br />

or by other means: The Nature Conservancy of <strong>Utah</strong>, US Forest Service Rocky Mountain Research Station, University<br />

of <strong>Utah</strong> Department of Biology, the Flora of North America project, Providia, <strong>Utah</strong> Natural History Museum,<br />

<strong>Utah</strong> Botanical Center, Red Butte Garden and Arboretum, the state of <strong>Utah</strong> Department of Natural Resources, and<br />

Bio-West, Inc.<br />

Twenty presenters at the conference kindly prepared manuscripts for this inaugural issue of Calochortiana, which<br />

will serve as the official proceedings document for the conference. Thanks to all the contributors for their willingness<br />

to share, and for their patience. - Walter Fertig<br />

Abstracts of Presentations and Posters<br />

not Submitted for the Proceedings<br />

Biogeography of the Intermountain Region and<br />

Connections to the Southwestern USA<br />

Noel H. Holmgren, Curator Emeritus, New York Botanical<br />

Garden.<br />

Abstract: The Intermountain Region lies between the<br />

east base of the Sierra-Cascade mountain chain and the<br />

west side of the Rocky Mountains. Its southern boundary<br />

overlooks the warm deserts of the Southwest and the<br />

northern boundary lies along the base of the Oregon and<br />

Idaho batholiths. The Great Basin occupies nearly twothirds<br />

of the Region, and the Wasatch and Uinta Mountains<br />

and the <strong>Utah</strong> segment of the Colorado Plateau occupy<br />

the eastern third. In combination, the plant associations,<br />

vegetation zones, and plant species distinguish the<br />

region as a reasonably natural floristic unit, but there are<br />

many geologic-historical relationships with the Southwest.<br />

With the changing climate, even greater similarities<br />

may be anticipated in the future. The basin and<br />

range topography of the Great Basin offers a perfect<br />

place to monitor possible species migration from south<br />

to north and from valley to mountain.<br />

Flora of the Arizona Strip<br />

Duane Atwood, Brigham Young University, retired<br />

(1870) C.C. Parry (1874-1875); and later by A.L. Siler<br />

and Marcus E. Jones. Generally speaking, most Arizona<br />

botanists have given little attention to this area. The first<br />

concentrated effort of collecting on the Strip was by<br />

Ralph K. Gierisch, a retired Forest Service employee,<br />

who worked primarily as a volunteer for the BLM Arizona<br />

Strip District located in St. George, <strong>Utah</strong>. Ralph<br />

made hundreds of collections, which are deposited at<br />

that office with many duplicates at BYU and NAU. My<br />

interest and first collection from the Strip was made 27<br />

May 1968 1 mile north of Fredonia to secure the type<br />

for Phacelia constancei, while working on a revision of<br />

the crenulatae group of Phacelia (Hydrophyllaceae).<br />

Then later in 1970 while living in Fredonia and working<br />

on the Kaiparowits Environmental Impact Studies with<br />

BYU, thru 1975; and as the first botanist for BLM and<br />

the second one nationally for the Cedar City BLM District<br />

(1975-1977). Collection trips to this unique area<br />

continued through to the present, often with Larry C.<br />

Higgins. An annotated list of vascular plants has been<br />

generated for the entire Strip and the National Parks and<br />

Monuments within its borders. Six new endemic taxa<br />

have been described from the area: Phacelia higginsii,<br />

P. furnisii, P. hughesii, Camissonia dominguezescalantorum,<br />

Physaria arizonica var. andrusensis and<br />

Tetradymia canescens var. thorneae.<br />

Abstract: The "Arizona Strip" is a unique botanical<br />

area isolated from the rest of Arizona by the Colorado<br />

River. Our knowledge of its flora has been slow and<br />

incremental with a few collections from the early botanists<br />

who visited southern <strong>Utah</strong> such as Edward Palmer<br />

4


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Biogeography and the Evolution of Rare, Endemic<br />

Species: Insights from two Mustard Genera in the<br />

Southwest (Draba and Boechera, Brassicaceae).<br />

Loreen Allphin, Department of <strong>Plant</strong> and Wildlife Sciences,<br />

Brigham Young University, Provo, UT and<br />

Michael D. Windham, Duke University, Durham, NC.<br />

Abstract: With a growing number of plant species in<br />

danger of extinction due to human induced threats,<br />

species-by-species approaches to management are becoming<br />

unrealistic. Conservation of rare plants would<br />

be improved by a clearer understanding of the evolutionary<br />

forces that give rise to rare, endemic species.<br />

These data might facilitate the development of management<br />

strategies applicable to a wide range of rare species.<br />

For this study we conducted a detailed survey of<br />

species within the genera Draba and Boechera from the<br />

Southwest United States (genera and a region with the<br />

high concentrations of endemic species). We collected<br />

data on geographic distribution, degree of endemism,<br />

chromosome number, ploidy level, breeding system,<br />

reproductive fitness and presumed mode of speciation.<br />

The study revealed some interesting evolutionary and<br />

biogeographic patterns. Some rare endemic species in<br />

these genera were primarily diploid, outcrossing, paleoendemic<br />

species with relatively low fecundity. Conversely,<br />

other endemic species in these groups were primarily<br />

polyploid, autogamous or apomictic, neoendemics<br />

with relatively high fecundity. These patterns appear<br />

to reflect both the type of speciation that occurred and<br />

the geologic/biogeographic history of the region. The<br />

geography of rarity and endemism in these genera appears<br />

to be an expression of primary divergence, reticulate<br />

evolution, and evolutionary time.<br />

Physical and Chemical Characteristics of Xeric Soils<br />

in Eastern Great Basin Determines the Natural <strong>Plant</strong><br />

Associations, but Recently Ruderal Species have Become<br />

an Important Factor.<br />

Rodd Hardy, Bureau of Land Management, Salt Lake<br />

City, UT<br />

Abstract: What are major physical and chemical properties<br />

of the soil profiles that are key factors for different<br />

plant associations? What are the major natural plant<br />

associations based on these soil properties? What particular<br />

ruderal species, to what degree, and when did<br />

these invaders become a major role within plant associations<br />

today? What recommendations are needed to mitigate<br />

the impacts of ruderal species to natural plant communities?<br />

This paper will note the ecotone sharpness in<br />

which plant communities in dry climates change from<br />

one type to another for major plant species and the<br />

chemical properties of the soil which determine specific<br />

plant communities. Winterfat and gray molly sites have<br />

particularly been vulnerable to annual grass and goosefoot<br />

forbs, but invasive species effects upon endemic<br />

species such as Pohl’s milkvetch and Small spring parsley<br />

has also been notable.<br />

Predictive Habitat Models for Arctomecon californica<br />

Torrey & Frémont and Eriogonum corymbosum<br />

Bentham var. nilesii Reveal for the Upper Las Vegas<br />

Wash Conservation Transfer Area, Nevada.<br />

Amy A. Croft, Thomas C. Edwards, Jr., Janis L. Boettinger,<br />

Glen Busch, James A. MacMahon, US Geological<br />

Survey and the Ecology Center, <strong>Utah</strong> State University<br />

Abstract: The Upper Las Vegas Wash Conservation<br />

Transfer Area (ULVWCTA), situated northwest of<br />

Las Vegas, Nevada provides habitat for two of the<br />

state’s special status species, Arctomecon californica<br />

Torrey and Frémont and Eriogonum corymbosum Bentham<br />

var. nilesii Reveal. In an effort to aid the Bureau of<br />

Land Management Las Vegas Field Office in conservation<br />

based decision making, we built a family of statistical<br />

models capable of predicting likely locations of each<br />

species in the ULVWCTA. To predict locations of the<br />

plant species, emphasis was placed on sensitivity, the<br />

ability of the models to predict where the species were<br />

located. A. californica sites were characterized by soils<br />

with low shear and compressive strength values, a low<br />

percentage of rock, and a physical soil crust. The most<br />

common soil type and vegetation association occupied<br />

by A. californica was the Las Vegas type (spring deposits)<br />

and the Ambosia dumosa-Atriplex confertifolia<br />

vegetation association. Models for A. californica had<br />

moderate to excellent predictive capabilities, with accuracies<br />

as reflected by sensitivity ranging from 75% to<br />

95%. Small sample sizes precluded construction of any<br />

models for E. corymbosum var. nilesii. Instead, we were<br />

able to successfully predict likely locations of E. corymbosum<br />

var. nilesii with the A. californica models. Overall,<br />

the models had predictive capabilities of sufficient<br />

accuracy to be used in conservation decisions for the<br />

ULVWCTA.<br />

Comprehensive Interactive <strong>Plant</strong> Keys for the Southwest<br />

Bruce S. Barnes, Flora ID Northwest, Pendleton, OR<br />

Abstract: <strong>Plant</strong> conservation and management for any<br />

given locality is a complex process which depends on<br />

reliable and continually updated information regarding<br />

what species are found and where. These critical data<br />

5


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

often require time-consuming initial and ongoing plant<br />

surveys. Computerized interactive keys produced over<br />

the past 14 years by the author greatly facilitate plant<br />

surveys by reducing the time to key unknown species by<br />

90% or more. This presentation will demonstrate the use<br />

of the plant identification software, to provide the audience<br />

with an understanding of the potential applications<br />

of this resource. The keys include all known vascular<br />

plants, both native and introduced, which grow outside<br />

of cultivation in 16 states and 4 Canadian provinces,<br />

including California, Nevada and <strong>Utah</strong>, with Arizona<br />

and New Mexico to be added in 2010. <strong>Plant</strong> characteristics<br />

may be selected in any order, with no forced<br />

choices. Terms are defined and illustrated, extensive<br />

references are included, and color photos are provided<br />

for over 99% of species. Synonyms and menus of genera<br />

and families are provided to reduce problems of<br />

changing nomenclature. The software is continuously<br />

updated with name changes, new plant finds, and new<br />

photos, with free annual updates available for purchasers.<br />

The keys are available by state or larger region. In<br />

most cases descriptive information is provided for separating<br />

subtaxa when present. These keys are a powerful,<br />

innovative tool to assist in providing timely plant survey<br />

data for plant conservation and management.<br />

The Climate Puzzle of Global Warming: It is not just<br />

about Chilies!<br />

Robert R. Gillies, Director/State Climatologist, <strong>Utah</strong><br />

Climate Center at <strong>Utah</strong> State University<br />

Abstract: In arid and semi-arid Western North America,<br />

observations of climate change point to an increase<br />

in average temperature that is greater than the rest of the<br />

world’s average. In line with such a warming trend in<br />

climate, several studies of the precipitation regime for<br />

the region have documented less snowfall as evidenced<br />

by decreases in snowpack as well as earlier snow melt,<br />

increased winter rain events and reduced summer flows.<br />

An ensemble of global climate model (GCM) projections<br />

for Western North America reflect just such conditions<br />

in that they suggest intensifying drying conditions<br />

to be the norm for the Southwest region due primarily to<br />

Hadley Cell intensification. Regions that lie to the<br />

Northwest, the GCMs have as benefiting from increased<br />

precipitation but in transitional zones, i.e., between the<br />

wetter and drier zones, any gains in projected precipitation<br />

are offset by the likelihood of an increased frequency<br />

of above normal temperatures during the summer<br />

months; such results suggest that an overall deficit<br />

in water resources is on the cards for much of the Intermountain<br />

West.<br />

Long-Term Perspectives on Vegetation: Paleoecology<br />

as a Tool for Conservation and Ecosystem Management<br />

Mitchell J. Power, <strong>Utah</strong> Museum of Natural History,<br />

Department of Geography, University of <strong>Utah</strong><br />

Abstract: Long-term studies on vegetation history have<br />

demonstrated the role of climate in controlling the<br />

composition and distribution of species through time.<br />

Paleoecological studies that use fossil plant and pollen<br />

offer many lessons from the past, including: 1) plant<br />

species respond individualistically to climate change, 2)<br />

vegetation composition during the last Ice Age, 21,000<br />

years ago, was very different than today, and 3) plants<br />

have “migrated” across hundreds to thousands of kilometers<br />

since the last Ice Age in response to changing<br />

climate and disturbance regimes. These lessons from<br />

paleoecology can be used to inform conservation efforts<br />

towards the protection of plant species that face unprecedented<br />

climate change. Traditionally, most “longterm”<br />

conservation studies span less than 50 years and<br />

therefore characterize historical variations in plant communities<br />

within a limited temporal domain. Conservation<br />

efforts aim to restore natural habitats and protect<br />

landscapes, but the question remains; what to restore<br />

things to? Through merging paleoecological knowledge<br />

with conservation objectives, land managers and conservationists<br />

are better positioned to make informed decisions<br />

to protect plant species as we experience the rapidly<br />

changing climate of the 21st century.<br />

The Southwest Region ‘GAP’ Program for Mapping<br />

Vegetation and Species<br />

Doug Ramsey, Department of Geography and Earth Resources,<br />

<strong>Utah</strong> State University, Logan, UT<br />

Abstract: The Southwest Regional Gap Analysis Project<br />

(SWReGAP) is an update of the Gap Analysis<br />

Program’s mapping and assessment of biodiversity for<br />

the five-state region encompassing Arizona, Colorado,<br />

Nevada, New Mexico, and <strong>Utah</strong>. It is a multi-institutional<br />

cooperative effort coordinated by the U.S. Geological<br />

Survey Gap Analysis Program. The primary objective<br />

of the update is to use a coordinated mapping<br />

approach to create detailed, seamless GIS maps of land<br />

cover, all native terrestrial vertebrate species, land stewardship,<br />

and management status, and to analyze this information<br />

to identify those biotic elements that are underrepresented<br />

on lands managed for their long term<br />

conservation or are ‘gaps.’<br />

6


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Responses of Colorado Plateau Drylands to Climate<br />

Change: Variability due to Land Use and Soil-<br />

Geomorphic Heterogeneity<br />

Mark E. Miller and Jayne Belnap, U.S. Geological Survey,<br />

Southwest Biological Science Center, Moab, UT<br />

Abstract: Dryland ecosystems comprise well over 50<br />

percent of the Colorado Plateau province and are subjected<br />

to land uses such as livestock grazing, recreation,<br />

and energy development. Low and variable amounts of<br />

precipitation constrain dryland resilience to land-use<br />

activities, making drylands particularly susceptible to<br />

persistent changes in structure, function, and capacity<br />

for providing key ecosystem services such as soil stabilization.<br />

Through multiple effects on soil and vegetation<br />

attributes, land use also mediates ecosystem responses<br />

to climate. Ecosystem responses to interactive effects of<br />

land use and climate vary spatially in relation to soil<br />

geomorphic properties such as texture, depth, horizonation,<br />

and topographic setting due to effects of these<br />

properties on water and nutrient availability, soil erodibility,<br />

and site susceptibility to hydrologic alteration by<br />

soil-surface disturbances. We use existing data from<br />

Colorado Plateau drylands to illustrate these concepts<br />

and to develop a set of testable hypotheses about climate-land-use<br />

interactions (i.e., how climate and land use<br />

each affect ecosystem resilience to the other) in relation<br />

to soil-geomorphic properties. For example, we predict<br />

that climate-land-use interactions in Colorado Plateau<br />

drylands will be greater on deep soils than on shallow,<br />

rocky soils because the former support grasslands and<br />

shrub steppe ecosystems that have been most extensively<br />

used and modified by livestock grazing. We also<br />

predict that climate-land-use interactions will be greater<br />

on relatively fine-textured soils than on coarse-textured<br />

soils because the former tend to be more susceptible to<br />

exotic plant invasions and hydrologic alteration following<br />

disturbance, and because they exhibit greater fluctuations<br />

in resource availability in response to precipitation<br />

variability. Variable ecosystem responses to climate<br />

due to land use and soil have implications for scientists’<br />

efforts to predict ecological consequences of climate<br />

change with sufficient detail to inform management decisions,<br />

and for decision makers’ efforts to prioritize and<br />

evaluate risks of different management strategies.<br />

Colorado Rare <strong>Plant</strong> Conservation Initiative, Saving<br />

Colorado’s Wildflowers<br />

tage by improving the stewardship of Colorado’s most<br />

imperiled plants. One hundred thirteen native plant species<br />

in Colorado are considered imperiled or critically<br />

imperiled by the Colorado Natural Heritage Program,<br />

meaning they are at significant risk of extinction. Of<br />

these species, 63 are endemic, growing only in Colorado<br />

and no place else in the world. Nearly 50% of our<br />

state’s imperiled native plants are considered poorly or<br />

weakly conserved. Unlike animals, Colorado has no<br />

state-level recognition or protection for plants. Impacts<br />

to Colorado’s rare plants are at an all-time high due to<br />

our rapidly expanding human population. Primary<br />

threats include habitat loss and fragmentation associated<br />

with resource extraction, motorized recreation, housing<br />

and urban development, and roads. Many rare plants are<br />

also at risk due to a simple lack of awareness regarding<br />

their precarious status. Despite the size and scale of<br />

these threats, we still have a chance to make a difference<br />

through strategic conservation actions, since healthy<br />

populations of many imperiled plants still exist. The<br />

goal of the Rare <strong>Plant</strong> Conservation Initiative is to conserve<br />

Colorado’s most imperiled native plants and their<br />

habitats through collaborative partnerships for the preservation<br />

of our natural heritage and the benefit of future<br />

generations.<br />

Rare <strong>Plant</strong> Management and BLM Policy<br />

Carol Spurrier, Bureau of Land Management, Washington,<br />

DC.<br />

Abstract: Rare plant conservation continues to be part<br />

of the multiple use mission of the Bureau of Land Management<br />

(BLM) in the United States. With continuing<br />

increases in the demand for all types of energy and other<br />

goods provided by the public lands, as well as landscape<br />

scale changes in natural vegetation due to increased<br />

wildfire and climate change, we wondered if the public<br />

lands that have been designated as part of the Natural<br />

Landscape Conservation System (NLCS) hold significance<br />

for protection of rare plant resources in BLM. We<br />

examined 2006 occurrence data on BLM lands from<br />

NatureServe within NLCS unit boundaries to determine<br />

rare plant species occurring in each unit. In this paper<br />

we discuss our findings for the different types of<br />

designations (wilderness, wilderness study areas, National<br />

Monuments and National Conservation Areas)<br />

within the System.<br />

Brian Kurzel, Colorado Natural Areas Program<br />

Abstract: The Colorado Rare <strong>Plant</strong> Conservation Initiative<br />

is a diverse partnership of public and private organizations<br />

dedicated to conserving our state’s natural heri-<br />

7


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Thinking and Acting Together to Preserve Uinta<br />

Basin Rare <strong>Plant</strong>s<br />

Joan Degiorgio, Northern Mountains Regional Director,<br />

The Nature Conservancy, Salt Lake City, UT<br />

Abstract: <strong>Utah</strong>’s Uinta Basin is home to dozens of endemic<br />

plant species. Nine of those species are at risk<br />

due to increased levels of energy development. Representatives<br />

from state, federal and local agencies, the Ute<br />

Tribe, private consultants, conservation groups, researchers,<br />

industry and others have come together as the<br />

Uinta Basin Rare <strong>Plant</strong> Forum to collectively address<br />

threats to these plants. Through a transparent, open<br />

planning process key ecological attributes of each plant<br />

were identified and rated for viability, stresses identified<br />

and specific strategies developed. This planning effort<br />

has been an excellent tool to “capture” the collective<br />

wisdom of local experts; and, with this knowledge, develop<br />

comprehensive strategies for the nine species simultaneously.<br />

With an agreed upon strategy, the Forum<br />

will work together implementing the highest leverage<br />

strategies and engaging more partners and dollars because<br />

of the solid foundation built through this planning<br />

process.<br />

Interagency Management Agreement for TES<br />

Species in Central <strong>Utah</strong><br />

David Tait, botanist, Fishlake National Forest, Richfield,<br />

UT.<br />

Abstract: The Bureau of Land Management Richfield<br />

Field Office, Fishlake National Forest, and Capitol Reef<br />

National Park share management responsibilities for<br />

many of the same Threatened, Endangered & Sensitive<br />

plant species (TES). To enable each of these agencies to<br />

better manage their shared TES species, they developed<br />

an interagency agreement in 1999 that enables them to<br />

pool their funding. This funding, which has been minimal<br />

at times, has been used to employ an interagency<br />

botanist and hire seasonals to survey and monitor these<br />

TES species throughout their ranges, regardless of<br />

agency boundaries. The BLM Price Field Office and the<br />

US Fish and Wildlife Service were added in 2007. This<br />

project has allowed us to: (1) conduct intensive surveys<br />

for target species on potential habitat within the project<br />

area, (2) determine potential for impact by visitor, recreational<br />

or livestock use on long-term viability of these<br />

rare plants, and (3) conduct long-term monitoring on<br />

several of the rare species. Between 1999 and 2008 approximately<br />

100,580 acres were surveyed for some 30<br />

TES plants species by the IA team. approximately<br />

27,500 acres on BLM, 37,920 acres on Capitol Reef,<br />

and 35,160 acres on lands administered by the Fishlake.<br />

Conservation Success for a Rare Idaho Endemic:<br />

Conservation Agreement and Botanical Special Interest<br />

Area for Christ’s Paintbrush<br />

Kim Pierson-Motychak, Sawtooth National Forest,<br />

Twin Falls, ID and Jeffrey E. Motychak, Motychak Environmental<br />

Consulting, Twin Falls, ID.<br />

Abstract: Christ’s Indian Paintbrush (Castilleja christii)<br />

is a rare species known from only one population in<br />

Southwestern Idaho, Cassia County. Due to its restricted<br />

distribution and vulnerability to threats, C. christii is<br />

designated as a Candidate species for Federal listing<br />

under the Endangered Species Act (ESA). Conservation<br />

Strategies were completed in 1995 and 2002 for the establishment<br />

of long-term monitoring protocols. These<br />

have been implemented from 1995 to present. In 2003,<br />

the portion of the population not included in the Mount<br />

Harrison Research Natural Area (RNA) was designated<br />

as the Mount Harrison Botanical Special Interest Area<br />

(BSIA). In 2005, a ten year Candidate Conservation<br />

Agreement (CCA) was signed between the US Fish and<br />

Wildlife Service and the USDA Forest Service. This<br />

CCA identified the key threats to the population which<br />

included: 1) non-native plant introduction and establishment,<br />

2) recreational impacts, 3) hybridization, 4) unauthorized<br />

livestock impacts, 5) road construction, maintenance,<br />

and facilities, and 6) natural threats. A total of 42<br />

conservation action items were committed to in the<br />

CCA. Results from the implementation of these conservation<br />

action items include aggressive non-native plant<br />

treatment, increased interpretive education, agency and<br />

public interaction, long-term demographic and reproductive<br />

monitoring, host-specificity determination, and<br />

preliminary pollination ecology. Population trends indicate<br />

that while plant densities within the communities<br />

have declined over the 13-year period, individual reproductive<br />

output (flowering stems/plant) has increased.<br />

Modeling Distributions of Rare <strong>Plant</strong>s in the Southern<br />

Great Basin of <strong>Utah</strong><br />

Marti Aitken, <strong>Utah</strong> State University and US Forest Service<br />

Pacific Northwest Research Station, Portland, OR;<br />

Leila M. Shultz, College of Natural Resources, <strong>Utah</strong><br />

State University, Logan, UT; and David W. Roberts,<br />

Department of Ecology, Montana State University,<br />

Bozeman, MT.<br />

Abstract: Field-validated landscape level predictive<br />

models identify potential plant habitat for rare plants in<br />

the Great Basin of western North America. Four rare<br />

species (Jamesia tetrapetala, Penstemon nanus, Primula<br />

domensis, and Sphaeralcea caespitosa) endemic to the<br />

southern portion of the eastern Great Basin (SW <strong>Utah</strong>)<br />

8


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

were chosen to include a range of environmental variability,<br />

growth form, and plant communities. Herbarium<br />

records of known occurrences were used to identify initial<br />

sample sites. We established multiple field sites to<br />

determine the geographic coordinates, environmental<br />

attributes (slope, aspect, soils, parent material) and<br />

vegetation data (associated species) in order to develop<br />

two predictive models for each species: a field key and a<br />

probability-of-occurrence or predictor map. The field<br />

key was developed from environmental attributes and<br />

associated species data collected at the sites and used<br />

only field data. Predictive maps were developed with a<br />

geographic information system (GIS) containing slope,<br />

elevation, aspect, soils, and geologic data – then randomly<br />

tested. Classification-tree (CT) software was<br />

used to generate dichotomous field keys and the maps of<br />

occurrence probabilities. Predictions from both models<br />

were randomly field-validated during the second phase<br />

of the study, and final models were developed through<br />

an iterative process. Data collected during the field validation<br />

were then incorporated into subsequent predictive<br />

models. The models identified potential habitat by<br />

combining elevation, slope, aspect, rock type, and geologic<br />

process into habitat models for each species. The<br />

cross-validated models were >96% accurate and generally<br />

predicted presence with accuracy >60%.<br />

Ute Ladies’-Tresses in the Diamond Fork Watershed:<br />

An Update<br />

Bridget M. Atkin and Steve R. Ripple, BIO-WEST,<br />

Logan, <strong>Utah</strong><br />

Abstract: Ute ladies’-tresses (Spiranthes diluvialis)<br />

(ULT) was listed as threatened in 1992. The largest<br />

known population is in the watershed of Diamond Fork<br />

Creek and its tributary, Sixth Water Creek. Between<br />

1916 and 2004, these streams were used as canals, and<br />

they conveyed irrigation water diverted from Strawberry<br />

Reservoir to the Wasatch Front. Increased peak flows<br />

altered the stream channel and aquatic ecosystem, creating<br />

unique conditions that allowed the rare orchid to<br />

thrive. In 2004 a system of pipes was installed to divert<br />

water directly into the Spanish Fork River, thereby reducing<br />

the flows in Diamond Fork and Sixth Water<br />

Creeks. Studies of ULT populations have been conducted<br />

since 1992 under the direction of <strong>Utah</strong> Reclamation<br />

Mitigation and Conservation Commission. Results<br />

show that ULT colonies are still maintaining large<br />

numbers. However, monitoring of ULT has been difficult.<br />

The unique life-cycle characteristics of ULT, along<br />

with its dynamic habitat, create many challenges.<br />

Highly variable yearly ULT counts are very difficult to<br />

interpret or correlate with environmental parameters. In<br />

2005 other studies were initiated and more associated<br />

plant species data are now being systematically collected<br />

to track changes that may indicate whether the<br />

decreased flows are impacting ULT habitat. During<br />

2007 ULT numbers showed at least two flushes, in early<br />

August with tiny individual plants. Conversely, in 2008<br />

ULT numbers were highest in mid-September and<br />

plants were large. These observations have wide reaching<br />

implications pertinent to many species, indicating<br />

that unless a population is observed carefully, data could<br />

easily be misinterpreted.<br />

Arizona Cliffrose (Purshia subintegra), An Arizona<br />

Endemic<br />

Debra Crisp, Coconino National Forest, Flagstaff, AZ,<br />

and Barbara G. Phillips, Zone Botanist, Coconino, Kaibab<br />

and Prescott National Forests<br />

Abstract: The Arizona cliffrose is a long-lived shrub,<br />

endemic to white Tertiary (Miocene and Pliocene)<br />

limestone lakebed deposits that are high in lithium, nitrates,<br />

and magnesium and is an Endangered species. It<br />

occurs in four disjunct populations spread across an area<br />

of approximately 200 miles in central Arizona. Threats<br />

to Arizona cliffrose include livestock grazing, mineral<br />

exploration, road and utility corridor development, offhighway<br />

vehicle use, urban development and drought.<br />

In this poster we summarize the results of some longterm<br />

monitoring transects initiated in 1987. These transects<br />

are in the Cottonwood population, which is<br />

thought to be the healthiest and contains the most diverse<br />

age structure of the four known populations. Data<br />

on these transects were collected three times, in 1987,<br />

1996 and in 2008.<br />

Demography and Pollination Biology of Graham's<br />

Penstemon (Penstemon grahamii), a Uinta Basin Endemic;<br />

5-year results.<br />

Rita [Dodge] Reisor and Wendy Yates, Red Butte Garden<br />

and Arboretum, University of <strong>Utah</strong>, Salt Lake City,<br />

<strong>Utah</strong><br />

Abstract: Penstemon grahamii is a Uinta Basin endemic<br />

which grows on oil-shale outcrops of the Green<br />

River Formation. Long-term monitoring plots were established<br />

for P. grahamii to collect basic life history<br />

data, study pollination biology, and survey critical habitat.<br />

Research was conducted over 5 years (2004 to 2008)<br />

during May – June, at the Blue Knoll/Seep Ridge and<br />

Buck Canyon population sites located on BLM land.<br />

Data gathered includes rosette diameter, number of inflorescences,<br />

inflorescence height, flowers per inflorescence,<br />

number of fruiting individuals, and herbivory.<br />

The breeding systems study used the following treat-<br />

9


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

ments: autogamy, geitonogamy, xenogamy, and vector<br />

pollination as a control group. Surveys were based on<br />

historic Element Occurrence (EO) reports and surrounding<br />

habitat. Demographic data suggest that P. grahamii<br />

population size has remained fairly stable over the study<br />

period. Annual survivorship rates range from 47 – 82%,<br />

and mortality ranging from 6 – 36%. Flowering events<br />

are highly variable annually, ranging from zero flowering<br />

plants in 2006, up to 44% in 2004. As expected for<br />

the breeding systems results, the vector (control) produced<br />

the most fruits, then xenogamy, geitonogamy,<br />

and autogamy with the least fruits produced. Survey<br />

results found existing populations at each historic EO<br />

visited, and expanded the range and size for some occurrences.<br />

Current threats include high rates of herbivory,<br />

habitat loss, and fragmentation due to oil and gas<br />

development. It is unknown how the reproductive success<br />

of P. grahamii may be influenced by other development<br />

related impacts such as dust production and pollinator<br />

disturbance.<br />

Micropropagation Studies in Astragalus holmgreniorum<br />

Aaron R. Fry, Brett A. McGowan, and Julianne<br />

Babaoka, Ally Bench, Renée Van Buren and Olga R.<br />

Kopp, Department of Biology, <strong>Utah</strong> Valley University,<br />

Orem, <strong>Utah</strong>, 84058<br />

Abstract: Astragalus holmgreniorum, a species endemic<br />

to the northern areas of the Mojave Desert is<br />

listed as a federally Endangered species. Threats to the<br />

species stem from habitat destruction arising primarily<br />

from commercial and residential development, overgrazing<br />

by livestock, recreational vehicles, and mining<br />

operations. In an attempt to develop a micropropagation<br />

technique aimed at aiding in recovery efforts for the<br />

species, we report successful induction of shoots from<br />

callus tissue. Explants were taken from leaves (abaxial<br />

and adaxial surfaces) and from petioles. These were incubated<br />

in MS medium amended with 2,-4 D and BA to<br />

induce callus formation. Murashige and Skoog medium<br />

amended with 7 mg/L of 2,-4 D and 2 mg/L of BA induced<br />

the formation of embryos and plantlets. Current<br />

work focuses on the effects of varying concentrations of<br />

NAA, IBA, and IAA on root formation. Following root<br />

induction, we plan to acclimatize plantlets by incubating<br />

them in potting soil. Ultimately, we hope that this research<br />

may be used to aid in recovery efforts of this species.<br />

Geochemical Analysis of Tuffaceous Outcrops Associated<br />

with the Narrow Endemic, Penstemon idahoensis<br />

Welsh & Atwood<br />

Paul R. Grossl, William A. Varga, and Richard M.<br />

Anderson, <strong>Utah</strong> State University, <strong>Plant</strong>, Soils, and Climate<br />

Department and <strong>Utah</strong> Botanical Center, <strong>Utah</strong> State<br />

University.<br />

Abstract: Idaho penstemon (Penstemon idahoensis<br />

Welsh & Atwood) is a Sawtooth National Forest Sensitive<br />

plant species narrowly endemic to the Goose Creek<br />

drainage in northern Box Elder County, <strong>Utah</strong>, and adjacent<br />

southern Cassia County, Idaho. Idaho penstemon is<br />

a short, glandular, perennial forb comprised of several<br />

stems which emerge from a semi-woody caudex and<br />

topped with showy, blue flowers. Its distribution is<br />

restricted to dry, light-colored, sparsely vegetated, tuffaceous<br />

outcrops of Tertiary Salt Lake Formation<br />

sediments. Botanists have long recognized the association<br />

of endemic, often rare, plant species with unusual<br />

soils. Edaphic endemism is prominent among those<br />

plant associations which include ultramafic bedrock,<br />

such as serpentine, or calcareous bedrock such as limestone,<br />

chalk, dolomite, or gypsum. Edaphic endemic<br />

species frequently arouse conservation concern due to<br />

their restricted distributions or small plant population<br />

sizes. The focus of this investigation considered the<br />

question whether Penstemon idahoensis utilizes unusual<br />

soil conditions that exclude other taxa and thus provide<br />

low competition environments. To answer this question<br />

we attempted to ascertain via geochemical analysis any<br />

selective or restrictive constituents or composition including<br />

the presence or absence of gypsum, unusual soil<br />

pH levels, soil texture, salinity (EC), organic matter, or<br />

atypical element distributions associated with common<br />

soil components.<br />

What's Happened to Siler Pincushion Cactus?<br />

Lee E Hughes, Ecologist, Arizona Strip Field Office, St.<br />

George, UT, retired<br />

Abstract: The Siler Pincushion Cactus has, like all<br />

vegetation in the southwest, been under the influence of<br />

a ten year drought. The effects of this drought are evident<br />

in the data gathered on the cactus. The poster will<br />

show the data from the six demographic plots on the<br />

cactus. The data starts in 1986 and goes through to<br />

2008. It summarizes the mortality data. The size structure<br />

for each plot is shown graphically to demonstrate<br />

the affect of the drought on the size composition of the<br />

cactus. In summary the drought has reduced the small<br />

cactus significantly. Also shown, is the effect (or no<br />

effect) from livestock and ATVs being present in the<br />

10


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

plots. The plot data shows a cactus with drought problems,<br />

but in some areas is doing well also.<br />

Clark County (Nevada) Rare <strong>Plant</strong> Modeling and<br />

Inventory<br />

Sonja R. Kokos, Clark County Desert Conservation Program,<br />

Las Vegas, NV; David W. Brickey and Larry R.<br />

Tinney, TerraSpectra Geomatics, Las Vegas, NV; and<br />

Analie R. Barnett and Robert D. Sutter, The Nature<br />

Conservancy, Southeastern Region, Durham, NC<br />

Abstract: To understand the distribution of rare plants<br />

covered under the Clark County Multiple Species Habitat<br />

Conservation Plan, Clark County and Terra-Spectra<br />

Geomatics developed two predictive GIS models. The<br />

models used ASTER Imagery and Landsat ETM+ Imagery,<br />

soils data (NRCS SSURGO), geologic data, and<br />

presence/absence data for eight rare and endemic plant<br />

species. The first model was used to predict the distribution<br />

of three gypsum loving species, the Las Vegas<br />

bearpoppy (Arctomecon californica), Sticky ringstem<br />

(Anulocaulis leiosolenus var. leiosolenus), and Las<br />

Vegas buckwheat (Eriogonum corymbosum var. nilesii).<br />

The second model was used to predict the distribution of<br />

five sand or potentially sand loving species, the Threecorner<br />

milkvetch (Astragalus geyeri var. triquetrus),<br />

Pahrump Valley buckwheat (Eriogonom bifurcatum),<br />

Sticky buckwheat (Eriogonum viscidulum), Beaver Dam<br />

breadroot (Pediomelum castoreum), and Whitemargined<br />

beardtongue (Penstemon albomarginatus).<br />

Using these models, Clark County can now describe the<br />

occurrence of all eight species in terms of high, medium<br />

and low probabilities. During the 2009 and 2010 field<br />

seasons, the county will test both models using a sampling<br />

protocol developed jointly by Clark County and<br />

The Nature Conservancy. Intuitively, we expect these<br />

models to predict the distribution of some species better<br />

than others, and further model refinement will be<br />

needed. However, the results to date have produced<br />

some interesting hypotheses regarding the life history<br />

and biology of these species. We expect the results will<br />

be valuable to Clark County and the federal land management<br />

agencies charged with managing these species.<br />

Post-Fire Monitoring of Erosion Resistance and Dust<br />

Emission on the Milford Flat Fire, West-Central<br />

<strong>Utah</strong><br />

Mark E. Miller, National Park Service, Moab, UT<br />

(formerly U.S. Geological Survey, Southwest Biological<br />

Science Center, Kanab, UT)<br />

Abstract: Soil stabilization is a major objective of postfire<br />

emergency stabilization and rehabilitation (ES&R)<br />

projects, yet monitoring data are rarely sufficient to determine<br />

whether treatments effectively achieve this objective.<br />

To address this information need, the U.S. Geological<br />

Survey and Bureau of Land Management are<br />

collaboratively monitoring effects of ES&R treatments<br />

on soil-surface stability and rates of dust emission in<br />

low-elevation portions of the 147,000-ha Milford Flat<br />

Fire that occurred in west-central <strong>Utah</strong> in July 2007. In<br />

August 2008, monitoring plots were established to<br />

evaluate the effectiveness of three types of ES&R treatments<br />

(aerial seeding and chaining, seeding with a<br />

rangeland drill, and seeding with a rangeland drill after<br />

herbicide application) in areas where field observations<br />

and satellite imagery indicated high rates of dust emission<br />

during spring 2008. Monitoring attributes include<br />

indicators of erosion resistance (soil stability, ground<br />

cover, and sizes of gaps between plant canopies) in addition<br />

to measures of plant cover and community composition.<br />

Seasonal rates of dust emission are currently<br />

monitored with BSNE dust samplers. Sampling in August<br />

2008 indicated that average soil-surface stability<br />

was highest in unburned control plots and in burned<br />

plots that were not treated. Average soil stability was<br />

lowest in burned plots that were seeded with a rangeland<br />

drill following herbicide application. During the August-October<br />

2008 period, rates of wind-driven soil<br />

movement varied over three orders of magnitude and<br />

were greatest in plots that received ESR treatments,<br />

were in exposed landscape settings, and had soils that<br />

were most susceptible to wind erosion.<br />

Using GIS and Remote Sensing to Predict Dominant<br />

<strong>Plant</strong> Species Distributions in Rich County, <strong>Utah</strong><br />

Kate Peterson, Doug Ramsey, Leila Shultz, John Lowry,<br />

Alexander Hernandez, and Lisa Langs-Stoner, Remote<br />

Sensing/GIS Laboratory and Floristics Lab, Dept. of<br />

Wildland Resources, <strong>Utah</strong> State University, Logan, UT<br />

Abstract: This research shows models of the potential<br />

spatial distribution of key upland plant species in Rich<br />

County, <strong>Utah</strong>. We used geospatial data layers of abiotic<br />

factors and remotely sensed (RS) imagery in conjunction<br />

with field-collected vegetation data. <strong>Plant</strong> species<br />

distribution maps are used to objectively and costeffectively<br />

correlate soil maps units with GIS data in the<br />

production of Ecological Site Descriptions (ESD’s).<br />

These were produced for Rich County in accordance<br />

with NRCS (Natural Resources Conservation Service)<br />

standards. Inasmuch as abiotic factors and vegetation<br />

associations can be used to predict the potential distribution<br />

of rare plants, we believe these analyses can be<br />

used to guide field searches for populations of endemic<br />

species.<br />

11


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

A Newly Discovered Gutierrezia on the Colorado<br />

Plateau<br />

Al Schneider, www.swcoloradowildflowers.com and<br />

Peggy Lyon, Colorado Natural Heritage Program<br />

Abstract: The authors will present information on the<br />

newly discovered species, Gutierrezia elegans or<br />

Lone Mesa Snakeweed. They discovered G. elegans<br />

August 4, 2008 while doing a plant survey in the new<br />

Lone Mesa State Park, 30 miles north of Dolores, Colorado.<br />

See http://www.swcoloradowildflowers.com/<br />

Yellow%20Enlarged%20Photo%20Pages/gutierrezia%<br />

20elegans.htm for details, including the full description<br />

published in the <strong>December</strong>, 2008 issue of the Journal of<br />

the Botanical Research Institute of Texas.<br />

Incorporting Demography, Genetics, and Cytology<br />

into Long-Term Management Plans for a Rare,<br />

Endemic Alpine Species: Draba asterophora<br />

Emily Smith and Loreen Allphin, Department of <strong>Plant</strong><br />

and Wildlife Sciences, Brigham Young University,<br />

Provo, UT.<br />

Abstract: Draba asterophora (Brassicaceae), a rare and<br />

endemic mustard, is known from three population clusters<br />

(North, Southeast, and south) occupying a narrow<br />

range of alpine habitats surrounding Lake Tahoe. The<br />

southern population cluster has been segregated as variety<br />

macrocarpa, whereas the other two clusters have<br />

been assigned to variety asterophora. Because this<br />

small, matted, perennial occurs at alpine sites, the species<br />

faces impending threats to its habitat through ski<br />

run expansion and development as well as from global<br />

climate change. With funding from the USDA Forest<br />

Service and local ski resorts, we are conducting morphological,<br />

ecological, chromosomal, and genetic studies<br />

of both varieties of D. asterophora to provide a<br />

framework upon which future management plans and<br />

mitigation can be developed. Preliminary results suggest<br />

that there are significant differences between the three<br />

population clusters. These include differences in soil<br />

composition, soil chemistry, plant density, demographics<br />

reproductive success, and genetics. Chromosome<br />

counts from the northern populations (Mt. Rose, Nevada)<br />

are tetraploid (n=20). Allozyme banding patterns<br />

support the hypothesis that these have arisen through<br />

autopolyploidy. The southeastern population has shown<br />

both diploid and triploid counts. Because the species<br />

includes more than one ploidy level, it should not be<br />

treated as a single panmictic taxon for purposes of conservation.<br />

Endangered Milkvetches of Washington County,<br />

<strong>Utah</strong><br />

Wendy Yates, Red Butte Garden and Arboretum, University<br />

of <strong>Utah</strong>, Salt Lake City, UT; and Ally Bench<br />

Searle and Renee Van Buren, <strong>Utah</strong> Valley State University,<br />

Orem, UT.<br />

Abstract: Astragalus ampullarioides (Welsh) Welsh<br />

and A. holmgreniorum Barneby are two federally listed<br />

Endangered species endemic to Washington County,<br />

<strong>Utah</strong>. A. ampullarioides is known from only four known<br />

populations, and A. holmgreniorum from only three<br />

populations. Prior research conducted on these species<br />

by Van Buren and Harper (2003) focused on vegetative<br />

and demographic characteristics. There have been no<br />

prior studies on the soil seed bank or seed viability. This<br />

study focused on determining the density of the soil<br />

seed bank and the percent seed viability for both species.<br />

For Astragalus ampullarioides soil was removed<br />

from two of the four known populations. For Astragalus<br />

holmgreniorum soil was removed from ten densely<br />

populated areas. Seeds were sifted from the soil and collected.<br />

The seeds extracted were then tested for viability<br />

by allowing them to germinate. Those seeds that did not<br />

germinate were further tested using the Tetrazolium test.<br />

This study found Astragalus ampullarioides had a soil<br />

seed bank density of 50 seeds/m 2 soil and viability was<br />

68.2%. Astragalus holmgreniorum had a soil seed bank<br />

density of 1.8 seeds/m 2 soil and viability was 87.7%.<br />

Collaborative Conservation for Washington County,<br />

<strong>Utah</strong>’s Federally-Listed <strong>Plant</strong>s<br />

Elaine York and Gen Green, The Nature Conservancy,<br />

Salt Lake City, UT.<br />

Abstract: Washington County, <strong>Utah</strong> is home to fourfederally<br />

listed plants: the Dwarf bear poppy (Arctomecon<br />

humilis), Siler pincushion cactus (Pediocactus<br />

sileri), Holmgren milkvetch (Astragalus holmgreniorum)<br />

and Shivwits milkvetch (Astragalus ampullarioides).<br />

Each faces pressing threats, especially habitat<br />

loss and degradation from urban development, invasive<br />

plants, and off-road vehicle use. Through U.S. Fish<br />

and Wildlife (USFWS) coordination and the efforts of<br />

partners, many conservation actions have been completed<br />

including land acquisition, habitat fencing, habitat<br />

restoration, the establishment of Areas of Critical<br />

Environmental Concern, seed germination and pollinator<br />

research, community education efforts, a habitat<br />

management endowment and more. Conservation actions<br />

have been implemented by USFWS, Bureau of<br />

Land Management, Dr. Renée Van Buren, Dr. Kimball<br />

Harper, Dr. Susan Meyer, U.S. Geological Survey,<br />

12


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Washington County, Red Cliffs Desert Reserve, the<br />

Shivwits Band of the Paiute Tribe, Zion National Park,<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>, <strong>Utah</strong> Natural Heritage Program,<br />

School and Institutional Trust Lands Administration,<br />

The Nature Conservancy, and more.<br />

Spatial Landscape Modeling: The Land Manager’s<br />

Tool Box<br />

Elaine York and Louis Provencher, The Nature Conservancy,<br />

Salt Lake City, UT.<br />

Abstract: Facilitated by The Nature Conservancy, the<br />

Spatial Landscape Modeling Project quantitatively modeled<br />

the reference and current conditions for seventeen<br />

major vegetation types in the Grouse Creek Mountains<br />

and Raft River Mountains, a 1.1 million acre landscape<br />

in northwest <strong>Utah</strong>. Partners – including <strong>Utah</strong> Partners<br />

for Conservation and Development, Bureau of Land<br />

Management, Sawtooth National Forest, <strong>Utah</strong> Division<br />

of Wildlife Resources, National Resources Conservation<br />

Service, Quality Resource Management, and private<br />

landowners – shared management data to explore effectiveness<br />

of current management and developed computer-generated<br />

alternative management scenarios to<br />

consider options for optimal land health. Cutting-edge<br />

technology from remote sensing, GIS analysis and partner-informed<br />

computer models produced a number of<br />

tools to assist land managers in their understanding of<br />

large-scale vegetation dynamics, long-term management<br />

options and the importance of management cooperation<br />

across land-ownership borders.<br />

An Update on Ecological Investigations of the<br />

Shivwits Milk-Vetch (Astragalus ampullarioides),<br />

Washington County, <strong>Utah</strong><br />

Mark E. Miller and Rebecca K. Mann, formerly US<br />

Geological Survey, Southwest Biological Science Center,<br />

Kanab, UT; Rebecca Lieberg, Cheryl Decker and,<br />

Kathy Davidson, Zion National Park, and Harland Goldstein<br />

and James D. Yount, U.S. Geological Survey,<br />

Earth Surface Processes Team, Denver, CO<br />

Abstract: The Shivwits milk-vetch (Astragalus ampullarioides)<br />

is one of four federally protected plant species<br />

restricted to particular geologic substrates at the edge of<br />

the Colorado Plateau and Mojave Desert in Washington<br />

County, <strong>Utah</strong>. Since 2006, the U.S. Geological Survey<br />

and National Park Service (Zion National Park, ZNP)<br />

have been studying this species in relation to geology<br />

and soils, herbivory, exotic plants, and mycorrhizal<br />

fungi. Habitat studies in 2006 documented the species<br />

on a new geologic substrate and across a broad range of<br />

soils, expanding the concept of potential habitat. Consumption<br />

of inflorescences by native herbivores reduced<br />

reproductive output in a ZNP subpopulation by 90% in<br />

2006 (low production year) and 75% in 2008 (high production<br />

year). Preliminary analyses indicate no significant<br />

effects of exotic red brome (Bromus rubens) biomass<br />

on growth or reproductive output of established<br />

milk-vetch plants in the same subpopulation during<br />

spring 2008. Effects of brome biomass on seedling recruitment<br />

remain unclear because low precipitation in<br />

2006 and 2007 prevented seed collection required for<br />

experimental studies. In 2007, median soil seed bank<br />

density in plots at ZNP was 45.7 seeds m 2 , with an extremely<br />

high density (2741 seeds m 2 ) in the plot with the<br />

sandiest soil. Coarse textured soils in this plot may reduce<br />

germination frequency, thereby resulting in longterm<br />

seed accumulation. Overall, results to date indicate<br />

that caging to exclude native herbivores may be the least<br />

expensive way to improve the viability of extant populations<br />

by enhancing reproductive output.<br />

Population Genetic Structure of an Endangered<br />

<strong>Utah</strong> Endemic Astragalus ampullarioides (Welsh)<br />

Welsh (Fabaceae)<br />

Jesse W. Breinholt, <strong>Utah</strong> Valley University, Orem, UT<br />

and Brigham Young University, Provo, UT; and Renee<br />

Van Buren, Olga R. Kopp, and Catherine L. Stephen,<br />

<strong>Utah</strong> Valley University, Orem, UT<br />

Abstract: The Shivwits milkvetch, Astragalus ampullarioides<br />

(Welsh) Welsh, is a perennial herbaceous plant<br />

in the family Fabaceae. This <strong>Utah</strong> edaphic endemic was<br />

federally listed as Endangered in 2001 because of high<br />

habitat specificity and low numbers of individuals and<br />

populations. All known occupied habitat for A. ampullarioides<br />

was designated as critical habitat by the US<br />

Fish and Wildlife Service in 2006. We used AFLP<br />

markers to assess genetic differentiation among the<br />

seven extant populations and quantify genetic diversity<br />

in each. Six different AFLP markers resulted in 217 unambiguous<br />

polymorphic loci. We used multiple methods<br />

to examine how population genetic structure in this species<br />

has changed over time. The genetic data indicate<br />

that, relatively recently, A. ampullarioides consisted of a<br />

single large contiguous genetic unit that fragmented<br />

over time into 3 genetic regions. These regions further<br />

fragmented and extant populations have differentiated<br />

through genetic drift. Populations exhibit low levels of<br />

gene flow, even between geographically close populations.<br />

We suggest plans for population establishment or<br />

augmentation carefully consider the genetic makeup of<br />

each of the extant populations.<br />

13


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Summary of Sclerocactus Monitoring in the Uinta<br />

Basin<br />

Maria Ulloa, formerly Bureau of Land Management,<br />

Richfield, UT<br />

Abstract: Sclerocactus brevispinus is a small barrel<br />

cactus endemic to the Pariette Draw and S. wetlandicus<br />

is a larger barrel cactus endemic to the Green River<br />

benches. Both geographic locations are in the Uinta Basin.<br />

These species were listed as Threatened with Sclerocactus<br />

glaucus. Currently the species are under review<br />

by USDI-FWS; the agency is working on its final<br />

ruling to make the taxonomic changes to separate these<br />

3 species. In May of 1997, 37 monitoring plots of S.<br />

brevispinus and S. wetlandicus were established in the<br />

Uinta Basin, including 9 plots with transplanted individuals.<br />

Ecosphere Environmental Service (Ecosphere)<br />

entered a Cooperative Agreement with the Vernal Field<br />

Office of the Bureau of Land Management (BLM) to<br />

study the genus Sclerocactus in the Pariette Drainage.<br />

All cacti within a 15m-radius of the plot center point<br />

were mapped and tagged. In 1998, the plots were read<br />

by Ecosphere. Funding to continue the monitoring was<br />

not allocated. During the winter of 2004, BLM decided<br />

to relocate the plots for the 2005 field season to see if<br />

the tagged cacti could be found. The BLM was successful<br />

at relocating the plots and the tagged cacti and decided<br />

to continue the monitoring for 4 years. In addition<br />

to finding the fate of the tagged cacti, all new individuals<br />

have been mapped and tagged. Other information<br />

collected has been number of flowers, number of capsules,<br />

and a small sampling of how many seeds per capsule.<br />

During the field season of 2008, a random sampling<br />

of the distance of cacti from the center of ant’s<br />

nests was measured to determine if ants influenced distribution<br />

and dispersal of seeds.<br />

Demographics of Sclerocactus Species in the Uintah<br />

Basin<br />

Lynda Sperry, SWCA Inc., Salt Lake City, UT<br />

Abstract: Botanical surveys associated with oil and gas<br />

development in the Uintah Basin provide large databases<br />

for listed plant species in compliance with the Endangered<br />

Species Act. We have extensively surveyed<br />

two threatened cacti species over the past three years. A<br />

total of 8,793 individuals were identified in 2008, 4,780<br />

were Sclerocactus wetlandicus, 3,663 were S. brevispinus,<br />

and 350 were identified as possible hybrids. We<br />

found the greatest concentration of S. brevispinus on<br />

north facing slopes (20%), followed by flat surfaces<br />

(15%), and the least amount on east facing slopes (6%).<br />

The aspect was not as significant for S. wetlandicus with<br />

18%, 17%, and 16% on north, south, and flat surfaces<br />

respectively. The aspect with the least number of S. wetlandicus<br />

was northwest-facing with 6% of the individuals<br />

surveyed. Both species were found more often in<br />

communities dominated by Atriplex, including A. confertifolia,<br />

A. canescens, or A. corrugata. Using the<br />

SSURGO data layer, S. wetlandicus was found more<br />

frequently on Motto-Rock outcrop complex, whereas S.<br />

brevispinus was more frequently found on Badland-<br />

Rock outcrop complex. Both soil types are dominated<br />

by clay and have high salinity ratings. Utilizing survey<br />

data collected as part of the permitting process for oil<br />

and gas development provides a unique opportunity to<br />

gain basic ecological and demographic information for<br />

federally listed species.<br />

Breeding System Characterization of a Threatened,<br />

Cliff Dwelling, Narrow Endemic Primula<br />

Jacob B. Davidson and Paul G. Wolf, Biology Department,<br />

<strong>Utah</strong> State University, Logan, UT.<br />

Abstract: The maguire primrose (Primula cusickiana<br />

var. maguirei or Primula maguirei) is a Threatened cliff<br />

dwelling endemic plant found only in northern <strong>Utah</strong>’s<br />

Logan Canyon. Although a small number of populations<br />

are close to one another, these populations have highly<br />

differentiated genetic structure. Maguire primrose, like<br />

most species in the genus Primula, exhibits reciprocal<br />

herkogamy in its morphology. Pin morphs have stigmas<br />

that extend to the opening of the corolla tube and anthers<br />

found near the bottom of the tube. Thrum morphs<br />

have stigmas found near the bottom of the corolla tube<br />

and anthers that are near the mouth of the corolla tube.<br />

Despite the spatial separation of anthers and stigmas,<br />

self fertilization has been observed for a number of Primula<br />

species. In the spring of 2008, I made hand pollinations<br />

using intramorph illegitimate outcrossings, legitimate<br />

intermorph outcrossings, and various autogamy<br />

and geitonogamy tests, while excluding pollinators. Resulting<br />

seed set was examined from each treatment. Several<br />

temperature and relative humidity monitors were<br />

placed near the plants, to see if environmental conditions<br />

affected hand pollination success. Temperature<br />

fluctuations at each study site ranged widely between<br />

freezing temperatures and 20ºC. We share our preliminary<br />

results from this initial field season here.<br />

14


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

The Taxonomic Distinctness of Eriogonum corymbosum<br />

var. nilesii<br />

Mark Ellis, Biology Department, <strong>Utah</strong> State University,<br />

Logan, UT.<br />

Abstract: We examined populations of perennial,<br />

shrubby buckwheats in the Eriogonum corymbosum<br />

complex and related Eriogonum species in the subgenus<br />

Eucycla, to assess genetic affiliations of the recently<br />

named variety E. corymbosum var. nilesii. We compared<br />

AFLP profiles and chloroplast DNA sequences of<br />

plants sampled from Colorado, <strong>Utah</strong>, Nevada, northern<br />

Arizona, and northern New Mexico. We found evidence<br />

of genetic cohesion among Nevada's Clark County<br />

populations as well as their genetic divergence from<br />

populations of other E. corymbosum varieties and<br />

Eriogonum species. The genetic component uncovered<br />

in this study supports the morphological findings upon<br />

which the nomenclatural change was based, attesting to<br />

the taxonomic distinctness of this biological entity.<br />

Drymocallis and Other Generic Segregates from<br />

Potentilla (Rosaceae)<br />

Barbara Ertter, UC Berkeley, Curator of Western North<br />

American Flora<br />

Abstract: Generic delimitation in tribe Potentilleae<br />

(Rosaceae) has historically vacillated between a<br />

broadly circumscribed Potentilla and recognition of<br />

various segregate genera. Recent convergence of morphological<br />

and molecular studies has shown that several<br />

segregates are in fact more closely related to Fragaria<br />

than to core Potentilla. These are accordingly treated as<br />

Comarum palustre, Dasiphora fruticosa, Sibbaldia procumbens,<br />

Sibbaldiopsis tridentata, and multiple species<br />

of Drymocallis in a pending volume of Flora of North<br />

America (FNA). The last genus includes Potentilla arguta,<br />

P. fissa, and P. glandulosa in North America, as<br />

well as 10-20 Eurasian species (e.g., D. rupestris). However,<br />

rather than simply transferring the existing subspecies<br />

or varieties of P. glandulosa into Drymocallis, a<br />

provisional revision was undertaken to more closely<br />

approximate the natural variation that occurs in western<br />

North America. As a result, 15 species of Drymocallis<br />

are recognized in FNA, some with additional varieties:<br />

D. arguta, D. arizonica, D. ashlandica, D. campanulata,<br />

D. convallaria, D. cuneifolia, D. deserertica, D. fissa,<br />

D. glabrata, D. glandulosa, D. hansenii, D. lactea, D.<br />

micropetala, D. pseudorupestris, and D. rhomboidea.<br />

Some species and varieties are newly described, and<br />

additional variation was noted as potentially deserving<br />

taxonomic recognition or conservation attention. This<br />

revision of Drymocallis acknowledges the existence of<br />

wide zones of intergradation and ambiguous populations,<br />

countered by the philosophy that conservation and<br />

other needs are poorly served by too broad taxonomic<br />

circumscriptions that gloss over valid components of<br />

biodiversity in an ecogeographic setting.<br />

Doing Adaptive Management: Improving the Application<br />

of Science to the Restoration of a Rare<br />

Tahoe <strong>Plant</strong><br />

Bruce Pavlik and Alison Stanton, BMP Ecosciences,<br />

South Lake Tahoe, CA<br />

Abstract: Tahoe yellow cress (Rorippa subumbellata),<br />

a plant endemic to the shores of Lake Tahoe, has been a<br />

candidate for protection under the Endangered Species<br />

Act since 1999. In 2002, a conservation strategy that<br />

described an adaptive management process for directing<br />

research, management, and restoration of the species<br />

was adopted by 13 signatory stakeholders. Although the<br />

implementation phase is at least four years from completion,<br />

we believe it provides an operative example of<br />

science-driven decision making. Specifically, we have<br />

found that implementation of adaptive management can<br />

be successful if: 1) the conceptual model of the adaptive<br />

management process is modified to include benefits to<br />

biological resources in situ, 2) all stakeholders are included<br />

upfront in the adaptive management working<br />

group to participate in the strategy and design of the<br />

whole program, 3) key management questions are used<br />

to focus data collection and identify essential management<br />

actions, and 4) information flow and the sequence<br />

of project stages (actions) are designed to facilitate<br />

stakeholder responses. In addition, the chance of success<br />

is greatly increased when agencies carefully choose target<br />

resources that meet several corollary requirements.<br />

A program of experimental reintroductions of Tahoe<br />

yellow cress from 2003 to 2006 not only produced a<br />

wealth of knowledge useful to managers, it also released<br />

1.5 million new seeds and 10,000 new plantlets into appropriate<br />

habitats around Lake Tahoe. Such tangible<br />

benefit to the species prompted the U.S. Fish and Wildlife<br />

Service to downgrade the priority status of the species<br />

under ESA.<br />

15


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

An Annotated List of the Endemic Species of<br />

Arizona<br />

Andrew Salywon, Wendy Hodgson, Desert Botanical<br />

Garden, Phoenix, AZ; Todd Ontl, Desert Botanical Garden<br />

and Department of Ecology and Evolutionary Biology,<br />

Iowa State University, Ames, IA; and Martin<br />

Wojciechowski, School of Life Sciences, Arizona State<br />

University, Tempe, AZ<br />

Abstract: In the Flora of Arizona, Kearney and Peebles<br />

(1964) estimated that roughly 5% (ca.164 spp.) of the<br />

flora is endemic to the state, and identified southern Arizona<br />

as harboring nearly double the number of endemic<br />

species compared to other parts of the state. However,<br />

no list of endemic taxa was provided. Therefore, in order<br />

to make meaningful comparisons of the endemic<br />

diversity with other states and to identify “hotspots” of<br />

endemicity within Arizona, we are compiling and annotating<br />

a list of the endemic plant taxa in Arizona. The<br />

annotations include taxonomic synonomy, publication<br />

and typification information in addition to distributional,<br />

ecological and evolutionary relationship data. Our working<br />

list is composed of ca. 250 taxa from 43 families<br />

and identifies the northern portion of the state (namely<br />

the Arizona Strip and the Grand Canyon) as harboring<br />

the highest percentage of endemics, in contrast to Kearney<br />

and Peebles analyses. It is hoped that insights into<br />

the relationships between geographical patterns and biological<br />

processes that can be gained from the list, including<br />

comparisons of the timing and mode of evolution<br />

of different groups. For example, Astragalus,<br />

Perityle, Agave, Eriogonum and Penstemon have been<br />

identified as the genera with the most endemic species.<br />

Not surprisingly these genera are composed of mostly<br />

annuals to short-lived perennials, with the exception of<br />

Agave, and are in groups that have undergone rapid and<br />

recent diversification in the Quaternary. In contrast, the<br />

woody endemics Berberis harrisoniana, Rhus kearneyi<br />

ssp. kearneyi, Sophora arizonica (= Calia) & Purshia<br />

subintegra are most likely of Tertiary origin and relictual.<br />

16


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Southwestern Ciénegas: Rare Habitats<br />

for Endangered Wetland <strong>Plant</strong>s<br />

Robert Sivinski,<br />

New Mexico Forestry Division, Santa Fe, NM, retired<br />

Abstract. Ciénega refugia for rare plants are medium to low elevation wet meadows characterized by stable springs<br />

and seeps in arid regions. Ciénega soils are usually alkaline and highly organic. Several southwestern plant species<br />

are confined to these habitats, making ciénega biotic communities distinct from other riverine or lentic wetlands in<br />

the region. A comprehensive inventory of southwestern ciénegas has not been completed; however, these habitats are<br />

clearly rare and diminishing in extent. Groundwater depletion, erosion, conversion to agriculture or aquaculture, abusive<br />

grazing, and exotic weeds threaten most remaining ciénega habitats and the rare species confined to them. Government<br />

and non-profit conservation agencies are attempting to restore and preserve a few remnant pieces of previously<br />

large ciénegas in Arizona, New Mexico, and Texas. Healthy ciénega habitats require active and continuous<br />

management at great effort and expense, especially for weed control.<br />

DEFINITION AND DESCRIPTION<br />

‘Ciénega’ is Spanish for a swamp, bog, or marsh. It<br />

is also spelled ‘ciénaga’ throughout much of the Spanish-speaking<br />

world – especially South America and the<br />

Caribbean. The ‘ciénega’ spelling is prevalent in the<br />

American southwest and often used in northern México.<br />

The origin of the word ‘ciénega’ is thought to be a contraction<br />

of the Spanish words ‘cien aguas’ meaning ‘a<br />

hundred waters or fountains’ (Crosswhite 1985). This is<br />

an allusion to springs, seeps and wet ground over a large<br />

area instead of a single pool, slough, or stream channel.<br />

Ciénegas gained acceptance as distinct climax communities<br />

of ecological significance when Hendrickson<br />

and Minckley (1985) made a thorough assessment of the<br />

ciénegas of southeastern Arizona. They defined the ciénega<br />

climax community as mid-elevation (1,000-2,000<br />

m) freshwater wetlands with permanently saturated,<br />

highly organic, reducing soils occupied by a lowgrowing<br />

herbaceous cover of mostly sedges and rushes.<br />

Few woody plants occur in the ciénega flora and often<br />

only as riparian tree species around the drier margins.<br />

Ciénegas occur in arid landscapes with high rates of<br />

evaporation, so the soils at the drying wetland margins<br />

usually have surface crusts of alkali or salts that are the<br />

deposited dissolved solids of evaporated or transpired<br />

soil solutions.<br />

Ciénega biotic communities of the southwestern<br />

United States and northern México are almost always<br />

features of springs and spring seeps (Brown 1982, Hendrickson<br />

and Minckley 1985, Dinerstein et al. 2000).<br />

Not all springs support ciénegas, but almost all ciénegas<br />

are supported by springs. These arid-land springs arise<br />

where stable aquifers intercept the ground surface in<br />

artesian basins or along geologic faults and fractures.<br />

They are generally not associated with fluctuating alluvial<br />

aquifers in channels that are flood-scoured, so are<br />

more likely to be found in the upper reaches of small<br />

drainages near geologic faults and igneous extrusions, in<br />

karst topography, and on gentle slopes where waterbearing<br />

strata have been exposed by river erosion or<br />

scarps. Size of individual ciénegas varies greatly from<br />

less than one acre to several hundred acres and is an expression<br />

of spring flow and topography.<br />

Ciénega vegetation is usually highly productive and<br />

dense. A list of plant species for southeastern Arizona<br />

ciénegas was assembled by Hendrickson and Minckley<br />

(1985). Milford and others (2001) and Sivinski and<br />

Bleakly (2004) produced lists of ciénega plants for the<br />

Rio Pecos Basin of eastern New Mexico. Most individual<br />

ciénegas have relatively low plant species diversity,<br />

but contribute a productive and rare subset of wetland<br />

species and habitats to an otherwise arid landscape. The<br />

most common ciénega plants of the southwestern region<br />

are the open water (when present) emergents of bulrush<br />

(Schoenoplectus spp.) and cattail (Typha spp.); sedges<br />

and rushes of water-saturated soils (Eleocharis spp.,<br />

Carex spp., Bolboschoenus maritimus (L.) Palla, Fimbristylis<br />

spp.); salt- and alkali-tolerant inland saltgrass<br />

(Distichlis spicata (L.) Greene), scratchgrass<br />

(Muhlenbergia aperifolia (Nees & Meyer) Parodi), and<br />

Mexican or Baltic rush (Juncus arcticus Willdenow<br />

vars. mexicanus (Willdenow) Balslev or balticus<br />

(Willdenow) Trautvetter) on seasonally saturated and<br />

sub-irrigated soils; and alkali sacaton (Sporobolus airoides<br />

(Torrey) Torrey) on the drier ciénega margins.<br />

Woody plants are usually not a significant part of ciénega<br />

vegetation cover (Figure 1), but patches of shrubby<br />

willows (Salix spp.) or willow-baccharis (Baccharis<br />

salicina Torrey & Gray) may occur and the drier ciénega<br />

margins will often have riparian trees such as cottonwood<br />

(Populus spp.), Arizona ash (Fraxinus velutina<br />

Torrey), and tree willows.<br />

17


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 1. Blue Hole Ciénega in Guadalupe County, New Mexico (September 2008) with Wright’s marsh thistle<br />

(right-center) and Pecos sunflower (distant yellow background).<br />

As climax communities associated with aridland<br />

springs of the southwest, ciénega-types of vegetation<br />

associations should be expanded to include more floristic<br />

regions and physical conditions than just the southeastern<br />

Arizona ciénegas described by Hendrickson and<br />

Minckley (1985). Ciénega synonyms in the Great Basin<br />

deserts, Sonoran and Chihuahuan deserts can sometimes<br />

include ‘vega’, ‘wet meadow’, ‘saltgrass meadow’,<br />

‘alkali meadow’ and even ‘oasis’. They are not necessarily<br />

confined to medium elevations and can also occur<br />

around desert springs at low elevations. Water coming<br />

from ciénega springs may be fresh or somewhat salty. In<br />

short, most ciénega-type habitats are the wet meadows<br />

that form around aridland springs and seeps.<br />

It is the relative permanence of the spring features<br />

that make many ciénega habitats biologically distinct<br />

from other types of wetland communities. Ciénegas are<br />

typically positioned in the upper reaches of small drainages<br />

or above river channels where they are protected<br />

from the scouring floods that frequently modify river<br />

marshes and floodplains. Ciénega spring flows may<br />

vary, but are less susceptible to the flooding and drying<br />

than playa basin wetlands during moist and arid cycles<br />

of the climate. Sediment cores from San Bernardino<br />

Ciénega in southeastern Arizona show wetland conditions<br />

for most of the last 7,000 years (Minckley and<br />

Brunelle 2007) and at Cuatro Ciénegas in Coahuila the<br />

fossil pollen in sediments indicate nearly identical ecological<br />

conditions for more than 30,000 years (Meyer<br />

1973). Such springs are refugia for species that may<br />

have been more widespread and common during wetter<br />

periods of the Quaternary. Several fish and invertebrate<br />

species are now confined to only one or a few aridland<br />

springs and the ciénegas they support are small remnants<br />

of stable habitat for some rare and endangered<br />

plants.<br />

LOSING (WET) GROUND<br />

The interaction of humans with aridland springs and<br />

ciénegas is a prehistoric tale of early and prolonged dependence<br />

(Hanes 2008; Rhea 2008) with a more recent<br />

history of almost universal destruction or diminution<br />

during the last two centuries (Unmack and Minckley<br />

2008). Hendrickson and Minckley (1985) documented<br />

the history and demise of many southeastern Arizona<br />

ciénegas – mostly by arroyo cutting that dropped spring<br />

aquifers below the ground surface. The tragic loss of<br />

most large springs and ciénegas by water withdrawals<br />

and aquifer depletion in the Chihuahuan desert of Trans-<br />

Pecos Texas is also well documented by Brune (1981),<br />

El-Hage and Moulton (1998), and Poole and Diamond<br />

(1993). Of all the southwestern states, the ciénegas of<br />

18


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

New Mexico are the least studied and documented in<br />

published literature. The following are a few examples<br />

of some historic and extant New Mexico ciénegas that I<br />

have personally studied.<br />

San Simon Ciénega was one of the wetland jewels in<br />

the crown of the southwest. It was a wet valley bottom<br />

about five miles long and a half mile broad that straddled<br />

the New Mexico/Arizona border in the upper-most<br />

reach of the San Simon Valley of Hidalgo County. The<br />

perennial spring-run creek had emergent marsh vegetation<br />

surrounded by wet meadow bordered with riparian<br />

woodland in sharp contrast to the adjacent Chihuahuan<br />

Desert scrub. When I first visited in 1994, San Simon<br />

Ciénega was dead and being covered with mesquite<br />

(Prosopis glandulosa Torrey). Many decadent cottonwood<br />

and willows trees still survived as reminders of<br />

the former wetland. The situation is the same today, but<br />

a little grimmer as these senescent tree remnants continue<br />

to decline.<br />

The lower part of San Simon Ciénega in Arizona was<br />

destroyed just before the turn of the twentieth century<br />

by regional overgrazing of cattle and an arroyo erosion<br />

cut that lowered the water table (Hendrickson and<br />

Minckley 1985). The arroyo headcut was arrested by a<br />

dam at the New Mexico border and seemed to spare the<br />

New Mexican part of the ciénega. Then irrigated cotton<br />

farming moved into the Arizona side of the valley and<br />

intercepted the spring aquifer emanating from the Chiricahua<br />

Mountains. An anonymous and undated report at<br />

the New Mexico Energy, Minerals and Natural Res-<br />

ources Department, Forestry Division documents most<br />

of the following events. The spring-run creek stopped<br />

flowing in 1952, shortly after irrigated agriculture<br />

started. When the Mexican duck was listed as an Endangered<br />

species, the New Mexico Game and Fish Department<br />

and federal Bureau of Land Management attempted<br />

to create some open-water nesting habitat in the<br />

still wet valley bottom by detonating powerful explosives.<br />

The resulting crater pools were not suitable nesting<br />

habitat, created habitat for weedy plants, and the<br />

valley bottom continued to dry from irrigated farming in<br />

the adjacent uplands. A nesting pond was excavated and<br />

frequently pumped full of water, but was eventually<br />

abandoned after the Mexican duck was removed from<br />

the list of Endangered species for taxonomic reasons.<br />

All permanent wetlands quickly disappeared from the<br />

valley.<br />

Another loss of desert springs and ciénegas occurred<br />

at a cluster of large springs near the dry mouth of the<br />

Rio Mimbres in Grant County. The fates of Apache<br />

Tejo Spring, Cold Spring, Kennecott Warm Spring, and<br />

Kennecott Cold Spring were to be completely captured<br />

by wells to supply water to the copper mill at Hurley in<br />

the early twentieth century. Walking from creosote desert<br />

into these former wetlands to find dusty gray organic<br />

soil supporting only clumpy alkali sacaton and<br />

surrounded by the decades-old carcasses of big cottonwood<br />

trees can only be described as shocking and depressing<br />

(Figure 2). Only two springs in the area remain<br />

wet to this day – Faywood Hotspring on its travertine<br />

Figure 2. Dead riparian woodland surrounding dry ciénega at former Cold Spring in Grant County, New Mexico<br />

(April 1993).<br />

19


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

hill with much reduced flow and no ciénega, and nearby<br />

Faywood Ciénega, which is supplemented and maintained<br />

by piped-in water from a distant upland spring.<br />

And what might we have lost from these wetlands? The<br />

type collection of Cleome sonorae A. Gray (= Cleome<br />

multicaulis A.P. de Candolle) was made “near the Mimbres”<br />

in 1851 and not seen again in the state. Perhaps<br />

one of these dead spring ciénegas was the habitat for<br />

this rare wetland species that is now likely extirpated<br />

from New Mexico.<br />

Most New Mexicans think of Lake Valley as an<br />

abandoned mining town in Sierra County, but the original<br />

Lake Valley is three miles north of the town at Berrenda<br />

Creek. An igneous extrusion across the broad<br />

Lake Valley segment backed-up a series of seasonal<br />

marshes and a permanent spring and seeps with ciénega<br />

vegetation. Berrenda Creek has only intermittent flow<br />

during wet seasons and storm events, but the runoff captured<br />

in Lake Valley sediments slowly discharged into a<br />

perennial spring run at the base of the valley. A diversion<br />

levee around Lake Valley was constructed about a<br />

century ago that dried the wetlands, which were converted<br />

to irrigated agriculture. The diversion outlet<br />

caused the lower spring run to erode into a deeply incised<br />

channel that still supports riparian woodland, but<br />

has lost its ciénega habitats.<br />

Some of the best remaining examples in New Mexico<br />

of ciénega habitats around aridland springs occur in<br />

the Rio Pecos valley on karst topographies near Santa<br />

Rosa (Guadalupe County) and Roswell (Chaves<br />

County). Even these are degraded and shrinking in size.<br />

Hundreds of acres of seeping ground with spring runs<br />

and sinkhole lakes have become surrounded by the City<br />

of Santa Rosa. About half of the original ciénega habitats<br />

have been damaged or destroyed by excavations for<br />

fish hatchery ponds or public fishing ponds; filling for<br />

sport fields, buildings, roads and parking lots; and dense<br />

infestation by Russian olive (Elaeagnus angustifolia L.)<br />

and salt cedar (Tamarix chinensis Loureiro) forests.<br />

A 25-mile stretch of Rio Pecos valley from Roswell<br />

south to Dexter has sinkhole lakes, resurgent creeks,<br />

spring runs, and seeps – sometimes with extensive ciénegas.<br />

Some of these have also been damaged by fish<br />

hatchery operations (Dexter National Fish Hatchery)<br />

and recreational development (Bottomless Lakes State<br />

Park and Roswell Country Club). Aquifer depletion below<br />

a municipal well completely dried a mile-long ciénega<br />

north of Dexter and other springs and seeps are<br />

likely impaired by the irrigated agriculture that dominates<br />

the landscape on the west side of the river. Fortunately,<br />

large seeps and spring runs from the highly alkaline<br />

Salt Creek aquifer support some ciénega habitats on<br />

Bitter Lake National Wildlife Refuge near Roswell. Unfortunately,<br />

the wildlife focus on this refuge has caused<br />

extensive damage to ciénega habitats with the numerous<br />

dikes, diversions and drains constructed to enhance fish<br />

and waterfowl habitats – although this is recently changing<br />

with more attention being paid to rare wetland<br />

plants.<br />

RARE PLANTS OF SOUTHWESTERN<br />

CIÉNEGAS<br />

While several animal species are endemic to particular<br />

aridland springs or areas of spring features, very few<br />

ciénega plants are so narrowly endemic. Some notable<br />

exceptions include Amargosa niterwort (Nitrophila mohavensis<br />

Munz & Roos), Ash Meadows mousetails<br />

(Ivesia kingii S. Watson var. eremica (Coville) Ertter),<br />

Ash Meadows gumplant (Grindelia fraxinopratensis<br />

Reveal & Beatley), and spring-loving centaury<br />

(Centaurium namophilum Reveal, Broome & Beatley),<br />

which are endemic to Mohave Desert springs around the<br />

Ash Meadows region of southwestern Nevada and adjacent<br />

California. Another very rare ciénega plant with a<br />

very small geographic distribution is reclusive lady's<br />

tresses orchid (Spiranthes delitescens Sheviak). This<br />

Endangered orchid is presently known from only four<br />

ciénegas in close proximity near the international border<br />

of southern Arizona (Coleman 2000).<br />

Most rare ciénega plants have very broad distributions<br />

of several hundred, or sometimes more than a<br />

thousand, miles in length. They are rare species because<br />

their ciénega habitats are very rare. Widespread wetland<br />

species usually do not get much attention from rare<br />

plant botanists because of multiple-state (or country)<br />

distributions and the difficulties of accessing a class of<br />

habitat that is predominantly on private property. A few<br />

widespread ciénega species, however, are starting to get<br />

some much needed scrutiny by botanists and land managers<br />

in southwestern states.<br />

Only three extant populations of Parish’s alkali grass<br />

(Puccinellia parishii A.S. Hitchcock) were known at the<br />

time this species was proposed for inclusion on the Endangered<br />

Species list in 1994. This annual grass occupies<br />

the highly alkaline soils of aridland springs and<br />

ciénegas. The proposal to list gave southwestern field<br />

botanists the incentive (funding) to search for new<br />

populations, which located or confirmed a total of 30<br />

sites at seeps and ciénegas – 17 in New Mexico, 11 in<br />

Arizona, 1 in eastern California, and 1 in southwestern<br />

Colorado (FR Vol. 63, No. 186, 51329-51332, 9/25/98).<br />

USDI-Fish and Wildlife Service withdrew the proposal<br />

to list in 1998, but New Mexico still lists this plant as<br />

state endangered because it is a wetland species with<br />

less than 100 acres of total known occupied habitat<br />

(New Mexico Rare <strong>Plant</strong> Technical Council 1999).<br />

The Pecos sunflower (Helianthus paradoxus Heiser)<br />

is another good example of a widespread ciénega plant<br />

that is a threatened species. It occupies only alkaline<br />

spring ciénegas from western Texas to west-central New<br />

20


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Mexico. The dramatic and almost complete demise of<br />

aridland ciénegas from aquifer depletion in the Chihuahuan<br />

Desert of Texas left only two populations of Pecos<br />

sunflower in a region that probably contained several.<br />

Some of the New Mexico populations are also damaged<br />

or threatened by aquifer depletion and nearly all are degraded<br />

by exotic tree infestations (USDI-Fish & Wildlife<br />

2008). Pecos sunflower was listed as a federally<br />

threatened species in 1999 and its ciénega habitats are<br />

finally receiving some management attention specific to<br />

the needs of this plant.<br />

Wright’s marsh thistle (Cirsium wrightii A. Gray)<br />

sometimes occurs in the same New Mexican ciénegas<br />

occupied by Pecos sunflower, but there appear to be<br />

fewer thistle populations in the United States. It ranges<br />

from southeastern New Mexico to southeastern Arizona<br />

and northern Chihuahua and Sonora. The type locality<br />

and single Arizona location at San Bernardino Ciénega<br />

has not been seen again since that ciénega was dried by<br />

down-cutting of the adjacent Black Draw. Some New<br />

Mexico populations at Lake Valley, Sacramento Mountain<br />

springs, and the City of Roswell (Country Club)<br />

have also been extirpated (New Mexico Rare <strong>Plant</strong><br />

Technical Council 1999, Sivinski 1995, 2005). This is<br />

clearly a threatened ciénega species in the United States;<br />

however, the status of this plant in México is unknown.<br />

A dismal trend of aridland spring loss in México (Contreras<br />

and Lozano 2002, Unmack and Minckley 2008)<br />

offers little hope that this species is faring better south<br />

of the border. Cirsium mohavense (Greene) Petrak<br />

(synonym = Cirsium virginense Welsh) is a related wetland<br />

thistle that may be occupying the same sinking ship<br />

in the Mojave Desert except that the Mojave thistle is<br />

not exclusively a ciénega plant and also occurs in some<br />

hanging garden and riparian habitats (<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong><br />

<strong>Society</strong> 2008).<br />

Leoncita false foxglove (Agalinus calycina Pennell)<br />

also co-occurs with Pecos sunflower and Wright’s<br />

marsh thistle in a ciénega at Bitter Lake National Wildlife<br />

Refuge in southeastern New Mexico. It is otherwise<br />

only known from an extant population at the Diamond<br />

Y Spring ciénega in western Texas, another historic and<br />

ambiguous collection in western Texas, and an historic<br />

collection in Coahuila (Poole et al. 2007). This is another<br />

species with almost no data available on its status<br />

in México. It seems to be exceedingly rare, but much<br />

additional research must be accomplished to support the<br />

initial appearance of rarity.<br />

Additional botanical surveys of all ciénegas in the<br />

southwestern United States and northern México will be<br />

needed to fully understand ciénega plant distributions<br />

and the threats to their habitats. Botanists should consult<br />

southwestern ichthyologists, herpetologists and aquatic<br />

invertebrate biologists who have been much more ag-<br />

gressive in locating and gaining access to aridland<br />

springs. They can help determine which springs support<br />

ciénega habitats and may already know many of the<br />

landowners.<br />

MANAGEMENT CHALLENGES<br />

Some remnant southwestern ciénegas have been acquired<br />

by federal and state governments and The Nature<br />

Conservancy as natural preserves or wildlife refuges.<br />

These have usually been protected because of the rare or<br />

endangered animals inhabiting the actual spring features,<br />

but the rare ciénega plants also need to be considered<br />

in preserve management. Ciénegas are productive<br />

and dynamic biotic communities that have attracted use<br />

by large herbivores for millions of years. A protective<br />

fence and hands-off approach for preserve management<br />

may only yield a ciénega that is overgrown, thatchy,<br />

drying, and pest-ridden (Kodric-Brown and Brown<br />

2007, Unmack and Minckley 2008). Needs for grazing<br />

or fire prescriptions, aquifer protection or restoration,<br />

and weed control calls for active management.<br />

Restoration and management of ciénegas affected by<br />

arroyo cuts that have lowered the potentiometric surface<br />

of adjacent springs and seeps will require the very difficult<br />

task of aggrading incised channels (Minckley and<br />

Brunelle 2007, Turner and Fonseca 2008). The ground<br />

water of a dead or damaged ciénega may still be close to<br />

the surface, but requires significant sedimentation and<br />

restoration of sheet flow to bring the potentiometric surface<br />

back to ground level and re-establish a “living” ciénega.<br />

On the other hand, former ciénegas supported by<br />

spring aquifers that have been depleted by groundwater<br />

pumping are unlikely to resume surface flow and become<br />

“living” again for the foreseeable future.<br />

Blue Hole Ciénega in Santa Rosa, New Mexico was<br />

purchased by the State Forestry Division’s Endangered<br />

<strong>Plant</strong> Program in 2005 to preserve critical habitat for the<br />

endangered Pecos sunflower and Wright’s marsh thistle.<br />

This 116-acre ciénega was about one-third infested with<br />

Russian olive trees (plus salt cedar to a lesser extent),<br />

suddenly ungrazed by livestock, and illustrative of some<br />

vegetation management challenges in a ciénega preserve<br />

(Figures 3 and 4).<br />

Weed tree control was an immediate concern because<br />

the entire ciénega was rapidly trending towards Russian<br />

olive woodland. Inmate work crews with chainsaws and<br />

backpack herbicide sprayers spent a total of 3,600 manhours<br />

cutting trees, spraying stumps, and broadcasting<br />

slash during the late summer and autumn months when<br />

the soil surface was dry over much of the ciénega. Winter<br />

to summer was an unsuitable period for weed control<br />

because effective herbicides could not be used while the<br />

soil surface was constantly wet or pooling water. The<br />

initial percent kill for tree stumps was about 80%.<br />

21


Figure 3. Work crew felling Russian olive trees at Blue<br />

Hole Ciénega (September 2007).<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

wide by October when a work crew spent another 1,000<br />

man-hours treating the resprouts with herbicide.<br />

After almost three years of weed tree control, controlled<br />

fire, and expenditure of $80,000, Blue Hole Ciénega<br />

might be restored to a point where a small crew<br />

can annually destroy new weed tree seedlings within a<br />

few days. The Albuquerque Chapter of the <strong>Native</strong> <strong>Plant</strong><br />

<strong>Society</strong> of New Mexico has volunteered to annually inspect<br />

the ciénega for other weed species that could arrive<br />

and become established, if not quickly detected.<br />

Accumulations of dead, thatchy, native vegetation will<br />

still have to be managed. Since high densities of Pecos<br />

sunflower are desirable at this preserve, intensive livestock<br />

grazing during the growing season is not an option.<br />

Therefore, controlled fires at suitable frequency<br />

(yet to be determined) will be needed for the foreseeable<br />

future.<br />

Most southwestern ciénegas are in private ownership<br />

because the spring features associated with them are<br />

valuable assets in an arid region. Restoration, protection<br />

and management of these rare and unique habitats are<br />

costly and require perpetual effort. Government programs<br />

that acquire ciénegas or assist landowners with<br />

their management are greatly needed in the southwestern<br />

states.<br />

When the slash had dried for three to five months,<br />

the entire ciénega was burned in <strong>December</strong>. Winter is<br />

the only feasible burn period because Pecos sunflowers<br />

set seed and die in October and the seeds begin to germinate<br />

in late February. The burn and mop-up took two<br />

days and was conducted by three State Forestry wildland<br />

fire crews, one USDA-Forest Service wildland fire<br />

crew and a pumper crew from the Santa Rosa Municipal<br />

Fire Department. The fire carried very well through the<br />

thick fine fuel of the grass and rush cover, but was generally<br />

not hot enough to consume tree slash more than<br />

one-inch in diameter.<br />

The bare, open ground of the burned ciénega created<br />

optimum habitat for germination and growth of annual<br />

ciénega plants such as Pecos sunflower and seepweed<br />

(Suaeda calceoliformis (Hooker) Moquin-Tandon)<br />

(Figures 5 and 6). Perennial ciénega plants quickly developed<br />

a lush growth that covered the charred one- to<br />

twelve-inch diameter tree slash lying on the ground.<br />

Most of the Wright’s marsh thistle rosettes that had been<br />

burned made new rosettes and many bolted flower<br />

stalks the following autumn. Unfortunately, about 20%<br />

of the weed tree stumps resprouted, and the still living<br />

roots around each dead stump sprouted one or more new<br />

weed tree saplings. These were three to four feet tall and<br />

Figure 4. Controlled burn of Blue Hole Ciénega<br />

(<strong>December</strong> 2007).<br />

22


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 5. Photo reference point at Blue Hole Ciénega before treatment (August 2006).<br />

Figure 6. Photo reference point at Blue Hole Ciénega after treatment (September 2008).<br />

23


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

LITERATURE CITED<br />

Brown, D.E. 1982. Biotic communities of the American<br />

Southwest – United States and México. Desert<br />

<strong>Plant</strong>s 4(1-4): 1-341.<br />

Brune, G. 1981. Springs of Texas. Vol. 1. Branch-<br />

Smith, Inc., Fort Worth, TX. 566 p.<br />

Coleman, R.A. 2000. Orchids at a range limit in Arizona<br />

and New Mexico. North American <strong>Native</strong> Orchid<br />

Journal 6(3): 193-200.<br />

Contreras-Balderas, S. and M. de L. Lozano-Vilano.<br />

1994. Water, endangered fishes, and development perspectives<br />

in arid lands of México. Conservation Biology<br />

8: 379-387.<br />

Crosswhite, F.S. 1985. Editorial introduction to Desert<br />

<strong>Plant</strong>s 6(3).<br />

Dinerstein, E., D. Olson, J. Atchley, C. Loucks, S.<br />

Contreras-Balderas, R. Abell, E. Iñigo, E. Enkerlin, C.<br />

Williams, and G. Castilleja. 2000. Ecoregion-based conservation<br />

in the Chihuahuan Desert: A biological assessment.<br />

World Wildlife Fund. Available: www.worldwild<br />

life.org/what/wherewework/chihuahuandesert/WWF<br />

Binaryitem2757.pdf [27 February 2009].<br />

El-Hage, A. and D.W. Moulton. 1998. Evaluation of<br />

selected natural resources in parts of Loving, Pecos,<br />

Reeves, Ward, and Winkler counties, Texas. Texas<br />

Parks & Wildlife, Resource Protection Division: Water<br />

Resources Team, Austin. 40 p.<br />

Hanes, C.V. 2008. Quaternary cauldron springs as<br />

paleoecological archives. In: L.E. Stevens and V.J. Meretsky<br />

(eds.) Aridland Springs in North America: Ecology<br />

and Conservation. The University of Arizona Press<br />

and The Arizona-Sonora Desert Museum, Tucson. pp.<br />

76-97.<br />

Hendrickson, D.A. and W.L. Minckley. 1985. Ciénegas<br />

– Vanishing climax communities of the American<br />

Southwest. Desert <strong>Plant</strong>s 6(3): 131-175.<br />

Kodric-Brown, A. and J.H. Brown. 2007. <strong>Native</strong><br />

fishes, exotic mammals, and the conservation of desert<br />

springs. Frontiers in Ecology and the Environment 5<br />

(10): 549-553. Available: www.frontiersinecology.org<br />

[27 February 2009].<br />

Meyer, E.R. 1973. Late-Quaternary paleoecology of<br />

the Cuatro Ciénegas Basin, Coahuila, México. Ecology<br />

54(5): 982-995.<br />

Milford, E., E. Muldavin, Y. Chavin and M. Freehling.<br />

2001. Spring vegetation and aquatic invertebrate<br />

survey 2000. Final Report to Bureau of Land Management,<br />

Roswell Field Office. 91 p. Available: http://<br />

nhnm.unm.edu/ [27 February 2009].<br />

Minckley, T.A. and A. Brunelle. 2007. Paleohydrology<br />

and growth of a desert ciénega. Journal of Arid Environments<br />

69: 420-431.<br />

New Mexico Rare <strong>Plant</strong> Technical Council. 1999.<br />

New Mexico Rare <strong>Plant</strong>s. Albuquerque, NM: New Mexico<br />

Rare <strong>Plant</strong>s Home Page. http://nmrareplants.unm.<br />

edu (Latest update: 22 January 2009) [27 February<br />

2009].<br />

Poole, J.M. and D.D. Diamond. 1993. Habitat characterization<br />

and subsequent searches for Helianthus paradoxus<br />

(puzzle sunflower). In: R. Sivinski and K. Lightfoot<br />

(tech. eds.) Southwestern rare and endangered plants. Proceedings<br />

of the Southwestern Rare and Endangered<br />

<strong>Plant</strong> Conference; March 30‐April 2, 1993; Santa Fe,<br />

NM. NM Forestry Division, Miscellaneous Publication<br />

No. 2, Santa Fe. p. 53-66.<br />

Poole, J.M., W.R. Carr, D.M. Price and J.R. Singhurst.<br />

2007. Rare <strong>Plant</strong>s of Texas. Texas A&M University Press,<br />

College Station. 640 p.<br />

Rhea, A.M. 2008. Historic and prehistoric ethnobiology<br />

of desert springs. In: L.E. Stevens and V.J. Meretsky<br />

(eds.) Aridland Springs in North America: Ecology and<br />

Conservation. The University of Arizona Press and The<br />

Arizona-Sonora Desert Museum, Tucson. p. 268-278.<br />

Sivinski, R.C. 1995. Wright’s marsh thistle (Cirsium<br />

wrightii). 1995 Progress Report (Section 6, Segment 10)<br />

to USFWS Region 2 Office, Albuquerque, NM. 8 p.<br />

Sivinski, R.C. 2005. Intermittent monitoring of Parish’s<br />

alkali grass (Puccinellia parishii), Wright’s marsh<br />

thistle (Cirsium wrightii), and Mescalero milkwort<br />

(Polygala rimulicola var. mescalerorum). Section 6,<br />

Segment 19 report to USFWS Region 2 Office, Albuquerque,<br />

NM. 6 p.<br />

Sivinski, R.C. and D. Bleakly. 2004. Vascular plants of<br />

some Santa Rosa wetlands, east-central New Mexico. The<br />

New Mexico Botanist 29: 1-5. Available: http://aces.<br />

nmsu.edu/academics/rangescienceherbarium/ [27 February<br />

2009].<br />

Turner, D. and J. Fonseca. 2008. What is a ciénega?<br />

(and why do we care?). Restoring Connections, Newsletter<br />

of the Sky Island Alliance 11(2): 1-4.<br />

Unmack, P.J. and W.L. Minckley. 2008. The demise<br />

of desert springs. In: L.E. Stevens and V.J. Meretsky<br />

(eds.) Aridland Springs in North America: Ecology and<br />

Conservation. The University of Arizona Press and The<br />

Arizona-Sonora Desert Museum, Tucson. p. 11-34.<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 2008. <strong>Utah</strong> Rare <strong>Plant</strong><br />

Guide. Salt Lake City, UT: <strong>Utah</strong> Rare <strong>Plant</strong> Guide<br />

Home Page. http://www.utahrareplants.org [27 February<br />

2009].<br />

USDI-Fish and Wildlife Service. 2008. Environmental<br />

assessment for designation of critical habitat for<br />

Pecos sunflower. Region 2, Albuquerque, NM. 64 p.<br />

24


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

A New Look at Ranking <strong>Plant</strong> Rarity for Conservation Purposes,<br />

With an Emphasis on the Flora of the American Southwest<br />

John R. Spence,<br />

Glen Canyon National Recreation Area, Page, AZ<br />

Abstract. A new rarity ranking system for prioritizing vascular plants for conservation and research is developed.<br />

This new system, termed the “At-Risk System” (ARS), ranks species using six variables, with each variable scored<br />

from 0-3; rarity type, biology (life-from, breeding system, pollination ecology, and dispersal ecology), population<br />

trend, anthropogenic threats, climate change vulnerability, and number of populations. Scores can range from 0 to<br />

18, with higher numbers indicating greater potential at-risk status. Selected species from the American Southwest are<br />

scored using the new system. Scores range from 2 for Pinus ponderosa to 18 for the critically endangered Arctomecon<br />

humilis. We know little about the biology and status of many rare species in the Southwest. This lack of<br />

knowledge was incorporated by using an automatic 3 score for the variables for which data are not available, which<br />

highlights uncertainty. For many species, the ARS score is not always strongly correlated with other ranking systems<br />

such as its ESA status or NatureServe G rank.<br />

The American Southwest supports one of the richest<br />

floras in North America, with perhaps as many as 6,000<br />

indigenous species distributed among the deserts and<br />

mountains of the region. The area includes six major<br />

arid and semi-arid biomes: the Chihuahuan, Colorado<br />

Plateau, Great Basin, Mohave, and Sonoran Deserts, and<br />

the Madrean region that extends from Mexico into<br />

southern New Mexico and Arizona. A recent compilation<br />

of rare species (Spence unpublished) has put the<br />

number of G1 and G2-ranked species at ca. 700. This<br />

preliminary list does not include the numerous local varieties<br />

of more common and widespread species, nor<br />

those species restricted to the Mexican portions of the<br />

biomes. With scarce resources, relatively few field botanists,<br />

and impending major climate change there is an<br />

urgent need to prioritize these species for conservation<br />

purposes. Unfortunately, very little is known about the<br />

basic ecological and biological characteristics of many<br />

of them. Thus, although there are ca. 260 G1 species in<br />

the American Southwest, we know relatively little about<br />

how to prioritize this list based on the likelihood of<br />

near-term extinction. Any attempt to rank large numbers<br />

of species for conservation funding must thus use a<br />

"triage" approach, based on general biological characteristics<br />

when more detailed quantitative data are not<br />

available. Rarity is also an elusive characteristic that<br />

may not be well reflected in its G rank, as it can change<br />

through time and space, and is not easily defined or<br />

quantified. Thus what defines rarity can vary across<br />

different scales (Harper 1981).<br />

There are three principal ranking systems that exist<br />

for rare plants, two of which have been applied locally.<br />

The California <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> ranking system has<br />

been developed for the flora of California (California<br />

<strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> 2001). Their system includes four<br />

lists indicating general status inside and outside of California,<br />

combined with a subjective 3-number code indicating<br />

local distribution, rarity and risks. These values<br />

are based on fairly detailed information, which is often<br />

not available for rare plants elsewhere in the Southwest.<br />

The California 1B list of rare endemic species includes<br />

over 1,000 taxa. The Nature Conservancy developed the<br />

G ranking system, which was then applied to state-level<br />

species lists throughout the West. However, the G ranking<br />

is a rather coarse-scale tool given the large number<br />

of species with either a G1 or G2 rank regionally. The<br />

third system, the IUCN threat categorization (Mace<br />

1994; Mace and Lande 1991), has not been applied to<br />

southwestern rare plants. Other more detailed systems<br />

have been developed (e.g., Bond 1994; Kwak and Bekker<br />

2006) but often require detailed information on genetics<br />

and estimates of pollen flow and fecundity. There<br />

remains a need among field biologists, land managers<br />

and conservation planners for a general system that can<br />

quickly rank and prioritize species within the G1-G2-G3<br />

levels.<br />

A significant amount of research has been done in<br />

attempting to correlate species extinction and rarity with<br />

biological traits (see reviews in Brigham and Schwartz<br />

2003; Krupnick and Kress 2005; Kunin and Gaston<br />

1997). In a seminal paper, Rabinowitz (1981) developed<br />

a rarity matrix that was based on geography, habitat specialization,<br />

and local population abundance. This was<br />

further discussed and applied by Kruckeberg and Rabinowitz<br />

(1985). However, this matrix approach to rarity<br />

has been little utilized in rare plant conservation planning.<br />

Kruckeberg and Rabinowitz suggested several<br />

avenues of research to pursue that may provide insight<br />

into rarity, including molecular/genetic, habitat, demographic<br />

and breeding system characteristics. More re-<br />

25


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

cently, Bevill and Louda (1999) examined the literature<br />

and documented 71 plant variables that had been used in<br />

the study of rarity and extinction. Many of these require<br />

detailed demographic or lab-based genetics research,<br />

and are thus impractical at a scale of the American<br />

Southwest in the near-term. Also, many reviews and<br />

studies have failed to find a strong link between specific<br />

variables and rarity (e.g., Aizen et al. 2002; Lavergne et<br />

al. 2004), although other studies have found such links<br />

(e.g., Sjöström and Gross. 2006). The general consensus<br />

is that demographic and breeding system data and<br />

molecular investigations overlaid with climate envelope<br />

analyses provide the best research approaches to understanding<br />

rarity (Brigham and Schwartz 2003; Gitzendanner<br />

and Soltis1999; Kunin and Schmida 1997). Beyond<br />

these more detailed techniques, some regional surveys<br />

have found traits that are correlated with rarity,<br />

including life-form and longevity (Harper 1979;<br />

Sjöström and Gross 2006;), breeding system (Sakai et<br />

al. 2002; Sjöström and Gross 2006; Sodhi et al. 2008;),<br />

pollination ecology (Robertson et al. 1999; Sakai et al.<br />

2002; Sodhi et al. 2008) and dispersal ecology (Bond<br />

1994).<br />

Any ranking system should be general enough to be<br />

widely applicable, but also provide the ability to rank<br />

species from potentially low to potentially high "at-risk"<br />

status. It should also be based on what is known about<br />

the biology of rarity. I have developed a preliminary<br />

ranking system, with an emphasis on the arid and semiarid<br />

flora of the American Southwest, based on six elements.<br />

This system should be viewed as a draft attempt<br />

to provide an early warning ranking for those species<br />

that potentially may be most at-risk in the Southwest.<br />

The time-frame is the next 50 years, when anthropogenic<br />

threats and climate change are likely to push many<br />

rare plants towards extinction. The elements are rarity<br />

type, biology, population trends, number of populations<br />

(TNC element occurrences), direct anthropogenic<br />

threats, and climate change vulnerability. Each of these<br />

is discussed below, with examples and ranking criteria<br />

in tabular form.<br />

ELEMENTS OF THE RANKING SYSTEM<br />

Each element is scored from 0 (low risk) to 3 (high<br />

risk). The six elements are rarity, biology, population<br />

trends, anthropogenic threats, climate change vulnerability,<br />

and G rank (here called the N rank, see below in<br />

Table 3). The final score is called the at-risk score<br />

(ARS) for each species. Each of the elements is described<br />

below.<br />

1. Rarity Type<br />

Table 1 lists the seven forms of rarity as defined by<br />

Rabinowitz (1981), scored either 0 or 1 for each level of<br />

the three categories. Each of the eight types is also<br />

coded from RI-RVIII. Since the three variables used in<br />

the system are subjective, they need to be defined in the<br />

local context of the American Southwest. Widespread<br />

species are defined as those that are typically found<br />

across one or more provinces (more than one subprovince)<br />

as defined by McLaughlin (2007). Thus a<br />

widespread species could occur in the Southwestern<br />

Region, Sonoran Province, and both the Sonoran and<br />

Mohavian subprovinces. Geographically local species<br />

are defined as those endemic to a single subprovince,<br />

such as the Colorado Plateau or Great Basin. Habitat<br />

specialists are defined as species that are generally restricted<br />

to one or a very few similar soils or substrates,<br />

while habitat generalists occur across a variety of substrates<br />

and soil types. Wetland species are also typically<br />

characterized as habitat specialists. Abundance at local<br />

sites (population or element occurrences) is more difficult<br />

to define, and remains subjective in my system.<br />

Many endemic species can be quite abundant locally (cf.<br />

Lesica et al. 2006; Spence unpublished data), while<br />

other species always seem to be uncommon or sparse<br />

where they are found. A species can have a final score<br />

of 0 for widespread common habitat generalists, to 3 for<br />

local sparse habitat specialists.<br />

2. Biology<br />

Detailed studies have shown that traits most closely<br />

related to rarity include low population size, demo-<br />

Table 1. The Rabinowitz rarity matrix (Rabinowitz 1981) scored from 0-3 for the ranking system.<br />

Abundance at<br />

Sites<br />

Geographic Range Widespread (0) Local (1)<br />

Habitat Specificity Generalist (0) Specialist (1) Generalist (0) Specialist (1)<br />

Common (0) RI 0 RII 1 RIII 1 RIV 2<br />

Sparse (1) RV 1 RVI 2 RVII 2 RVIII 3<br />

26


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

graphic factors, competitive abilities, inbreeding depression,<br />

low pollen/ovule ratios and pollen limitation, or<br />

highly specialized breeding, pollination and dispersal<br />

systems (Brigham and Schwartz 2003; Byers and<br />

Meagher 1997; Cole 2003; Gitzendanner and Soltis<br />

2000; Purdy et al. 1994; Schemske et al. 1994; Walck et<br />

al. 1999). In some cases rarity may also occur due to<br />

extreme and rapid habitat loss in formerly common species<br />

(eg., Ge et al. 1999). In the absence of data on these<br />

variables, proxies are needed that can be linked to the<br />

biological characteristics of rare species. Although the<br />

literature is somewhat ambivalent (e.g., Aizen et al.<br />

2002; Bevill and Louda 1999; Sakai et al 2002), some<br />

studies have suggested that a variety of basic plant characteristics<br />

may be used to provide preliminary indications<br />

of possible vulnerability and general extinction<br />

probabilities.<br />

Thus self-incompatible breeders, species with highly<br />

specialized and rare pollinators, and species that depend<br />

on highly specialized dispersers are often considered to<br />

be more at risk than species with more generalized biology<br />

(Aizen et al. 2002; Bond 1994; Buchmann and Nabhan<br />

1996; Kwak and Bekker 2006). I have developed a<br />

set of four proxy variables based on the literature, with<br />

the caveat that these are only general indicators of potential<br />

at-risk status. Detailed studies are clearly needed<br />

for most species in the Southwest in order to determine<br />

the exact causes of rarity, which are likely to be taxon<br />

(genus, species) specific (Bevill and Louda 1999). The<br />

four proxy variables (traits) are life-form, breeding system,<br />

pollination ecology, and dispersal ecology. Each of<br />

these is discussed with examples in Table 2. This element<br />

is scored for a species by selecting the single highest<br />

score among the four biological traits, rather than<br />

averaging them.<br />

Table 2. The biology element scored using four principal biological traits to characterize the vulnerability<br />

of a species to extinction.<br />

Biological Traits Ranked Explanation Score Examples<br />

Long-lived woody species<br />

Short-lived woody species or<br />

long-lived herbaceous species<br />

Short-lived perennial<br />

herbaceous species<br />

Annual or biennial<br />

Life-form and longevity<br />

Long life buffers against environmental<br />

change (>100 yrs)<br />

Some vulnerability, but generation times<br />

tend to buffer against short-term<br />

changes (>25 yrs)<br />

Vulnerable, short generation time may<br />

not be able to track environmental<br />

changes (3-25 yrs)<br />

Extremely vulnerable, with seed bank<br />

longevity a critical factor in persistence<br />

of populations<br />

Breeding System<br />

0 Conifers, Coleogyne<br />

1 Atriplex, Ericameria,<br />

Pediocactus<br />

2 Astragalus, Eriogonum<br />

Penstemon<br />

3 Cryptantha, Ipomopsis,<br />

Phacelia<br />

Autogamous<br />

Mixed mating<br />

Facultative outcrossing;<br />

some autogamy<br />

Obligate outcrossing; xenogamy,<br />

dioecy, loss of sexual reproduction,<br />

or breeding system<br />

unknown<br />

Can set seed despite small numbers of<br />

individuals through selfing<br />

Flexible system that allows for reproduction<br />

with or without pollinators, selfing<br />

occurs<br />

Generally requires pollen transfer between<br />

individuals, with low selfing rates<br />

often associated with reduced fitness<br />

Requires more than one individual, specialized<br />

pollen transfer<br />

27<br />

0 Many annuals<br />

1 Many generalized insect<br />

pollinated taxa<br />

2 Many insect pollinated taxa<br />

3 Orchidaceae, dioecious species


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 2. continued<br />

Biological Traits Ranked Explanation Score Examples<br />

Wind, water, self-pollination<br />

Generalized insects and vertebrates<br />

Generalized vertebrates<br />

Specialized animal pollination,<br />

or pollination system unknown<br />

Wind, water<br />

Generalized animal dispersal<br />

No structural or specialized features<br />

Specialized animal dispersal<br />

Pollination Ecology<br />

No requirements for animal vectors, can<br />

pollinate and set seed regularly<br />

Generally can pollinate and set seed<br />

through visitation of many different animal<br />

groups<br />

Generally can pollinate and set seed as<br />

long as hummingbirds, or other bird or<br />

bat species are present<br />

Requires a specialized, and often rare,<br />

insect or vertebrate species for pollination<br />

and seed set<br />

Dispersal Ecology<br />

Can be transported easily by wind or<br />

water<br />

Many bird and mammal species can<br />

disperse fruits<br />

No requirements for animal-mediated<br />

dispersal<br />

Requires a particular specialized animal<br />

vector to disperse seeds, such as Clark's<br />

nutcracker, Red crossbill, some antdispersed<br />

species<br />

0 Betulaceae, Conifers, Populus,<br />

aquatic species, many annuals<br />

1 Asteraceae, other open "dinnerplate"<br />

flowers<br />

2 Bat and hummingbirdpollinated<br />

species<br />

3 Agavaceae, Asclepias,<br />

Orchidaceae<br />

0 Spores, dust seeds, wind dispersed<br />

species<br />

1 Berry-producing species, sticktights,<br />

etc.<br />

2 Smooth seeds, short-distance<br />

dispersal only<br />

3 Fruits/seeds designed for specialized<br />

dispersal; Pinus albicaulis,<br />

Viola sp.<br />

3. Population Trends<br />

Mace (1994) suggested that population declines or<br />

loss of populations could be used to rank rare species.<br />

When data are available, this element can provide an<br />

accurate and straight-forward determination of the atrisk<br />

status of a species, but detailed information on<br />

population size and demographic data are difficult to<br />

collect and such data sets remain rare. General anecdotal<br />

information can be of some use in scoring a species<br />

in this element, however, such as loss of populations or<br />

general observational data showing declining numbers<br />

in populations. It is recommended that if time series,<br />

demographic, or abundance data are not available, that a<br />

species should be scored conservatively (higher), including<br />

a 3 when relevant as "unknown trends" (Table<br />

3).<br />

4. Direct Anthropogenic Threats<br />

Anthropogenic threats can include direct and indirect<br />

threats through climate change. I have separated out<br />

climate change from other more direct anthropogenic<br />

threats as the specific characteristics used to score them<br />

are different. Direct threats include water diversion or<br />

ground water pumping, recreational activity (e.g., OHV<br />

use), domestic livestock grazing, mining, agricultural<br />

development, introduction of invasive exotics, or urbanization.<br />

Threat levels can vary from minimal such<br />

as is found in many protected areas, to severe on lands<br />

around rapidly growing urban centers. This element is<br />

scored based on approximate values for impacts to habitat,<br />

ranging from generally minimal to >90% of a species<br />

habitat impacted. GIS analysis with ground surveys<br />

are often necessary to quantify the extent of impacts, but<br />

generally a rough first approximation can be made<br />

based on the knowledge of researchers familiar with the<br />

species and its habitat and associated threats. Thus the<br />

species scores remain somewhat subjective, based primarily<br />

on levels of degradation or impacts to the species<br />

overall habitat. Often this is not well known or is difficult<br />

to categorize. Especially difficult to quantify is the<br />

magnitude and imminence of these threats, and more<br />

work is needed to develop a more objective set of criteria.<br />

For now, the element is scored based on the presence<br />

of impacts regardless of intensity and timing. Gen-<br />

28


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

erally, when threats are poorly known and hard to quantify<br />

given available information, it is recommended that<br />

a conservative (higher) score be used. This element is<br />

scored range wide in this preliminary analysis (Table 3).<br />

5. Climate Change Vulnerability<br />

The climate envelope of a species can be modeled to<br />

predict distributional changes with global warming (e.g.,<br />

Miles et al. 2004). Although we lack data for many species,<br />

a proxy variable that can provide a first approximation<br />

to the vulnerability of a species to climate change is<br />

its elevation range. If a species has a relatively broad<br />

elevation range, populations along the gradient will be<br />

adapted to local climates, and the species genetic ability<br />

to adapt will likely be greater than in a species with a<br />

very small elevation range exposed to a much smaller<br />

range of climate variability. Elevation ranges from<br />

1000 meters are used to score each species<br />

for climate change vulnerability (Table 3).<br />

6. <strong>Number</strong> of Known Populations<br />

This element in the ranking system is based on the<br />

general concept of element occurrences developed by<br />

the Nature Conservancy and further refined by Nature-<br />

Serve (Faber-Langendoen et al. 2009). The five categories<br />

are the same as the G and S ranks on the various<br />

state and state natural heritage databases, although I<br />

have provided more explicit criteria for the G4 and G5<br />

levels. Species with widespread continuous populations,<br />

such as many conifers, are ranked G4 if found within a<br />

single subprovince, and G5 if they are in more subprovinces.<br />

Thus a widespread species such as Saguaro<br />

(Carnegiea gigantea) is ranked as a G4, despite its large<br />

population size, because of its restriction to the Sonoran<br />

Desert. Since there are slight differences between this<br />

ranking element and the G scale, I have given this element<br />

the letter N for number of known populations. The<br />

scores are inverted from the N rank, thus a N1 is scored<br />

a three, an N2 a two, a N3 a one, and N4N5 a zero.<br />

Table 3. Four additional elements of the scoring system for ranking at-risk species, with scores from 0-3,<br />

and explanations for how each is scored.<br />

Ranking Elements 3-6 Score Explanation<br />

3. Population Trend Historical trends (if known) since approximately 1900, or when the species was<br />

first discovered.<br />

Stable or increasing 0 Populations are secure for the foreseeable future, with trends exhibiting natural<br />

short and long-term variability<br />

Minor declines 1 Data are available to indicate that some declines have occurred, (10% decline in total numbers<br />

per year<br />

4. Direct Anthropogenic<br />

Threats<br />

Scores are typically applied range-wide, although they could be used in a more<br />

narrowly defined region if needed.<br />

No direct threats 0 Few if any impacts from recreation or domestic livestock grazing, no invasive exotics<br />

present, no mining activity, recreational impacts largely absent, etc.;


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 3. continued.<br />

Ranking Elements 3-6 Score Explanation<br />

5. Climate Change<br />

Vulnerability<br />

6. <strong>Number</strong> of known<br />

populations<br />

Estimate the elevational range based on population locality information.<br />

Low 0 Climate envelope and/or elevation range wide (>1000 m), apparently secure from<br />

climate change for the next 50 yrs<br />

Moderate 1 Climate envelope and/or elevation range moderate (>500 m), some vulnerability<br />

to climate change with moderate declines expected in next 50 yrs, secure for next<br />

10 yrs<br />

High 2 Climate envelope and/or elevation range small (100-500 m), vulnerable to climate<br />

change in the next 10 yrs, long-term (>50 yrs) probability of survival moderately<br />

low, with major population declines<br />

Severe 3 Climate envelope/elevation range very small (50 yrs) very low to none, declines<br />

or local extirpation expected within 10 yrs; or elevation range unknown; or<br />

an obligate wetland species in American Southwest<br />

N0 -- Extinct, at least in wild<br />

N1 3 1-5 known populations, endangered (=G1)<br />

N2 2 6-20 known populations, threatened (=G2)<br />

N3 1 21-100 known populations, vulnerable (=G3)<br />

N4 0 101-300 known populations, or relatively widespread and in more or less continuous<br />

stands, apparently secure (=G4)<br />

N5 0 >300 known populations or widespread and continuous, secure (=G5)<br />

For many species the definition of what constitutes a<br />

discrete population is difficult. In the case of widespread<br />

common species that occur in large continuous<br />

stands, the need to define discrete populations is of less<br />

importance than for rare species with localized scattered<br />

distributions In most cases, G1, G2, and G3 species<br />

tend to occur in scattered populations with gaps, where<br />

they are absent even in suitable habitat. For many of<br />

these species, gene flow and dispersal are likely to be<br />

relatively low, thus isolated populations, even if only a<br />

few hundreds of meters apart, can be effectively considered<br />

discrete and non-interacting populations. In Table<br />

3, those species with larger more continuous species<br />

(G4 or G5) can be scored without reference to the number<br />

of discrete populations.<br />

SCORING SPECIES USING THE SYSTEM<br />

For each species, the at-risk score is the sum of the<br />

scores for each of the six elements: rarity, biology, declines,<br />

anthropogenic threats, climate change vulnerabil-<br />

ity and N populations (see Tables 1-3), thus:<br />

At-risk Score (ARS) = rarity type + biology+ declines +<br />

threats + climate change + N score<br />

A local sparse habitat specialist with only 1-5 known<br />

occurrences would be scored very high. The larger the<br />

score, the more potentially at-risk the species should be<br />

considered as a first approximation. The overall conservation<br />

score can vary from 0 to 18. Examples for selected<br />

species are given in Tables 4 and 5 and are discussed<br />

below.<br />

General information on the status of a species should<br />

always be included for conservation planning (see Table<br />

5). Species should be characterized as part of the ranking<br />

system by at least some basic characteristics such as<br />

life-form, nativity, geographic distribution and protection<br />

status. Nativity would include three categories: indigenous<br />

(native, widespread), endemic (native, restricted<br />

to a subprovince), and paleoendemic (local, may<br />

30


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

be in one or more subprovinces, phylogenetically and<br />

geographically isolated; sensu Stebbins and Major<br />

1965). Geographic distribution is best characterized by<br />

the classification of McLaughlin (2007), who has developed<br />

the most current and detailed analysis of species<br />

distributions and floristic regionalization for the American<br />

Southwest. Protection status (land management<br />

categories) range from highly protected NGO preserves,<br />

many national parks, etc., through general public lands<br />

including wilderness, to state lands, and finally private<br />

lands. The categorization of Scott et al. (1993) provides<br />

a useful approach to categorize species. The percentage<br />

of the total number of populations in each of the land<br />

management categories should be determined, as it will<br />

provide useful additional information to make informed<br />

decisions about which species to prioritize for conservation<br />

funding.<br />

RESULTS<br />

If the system is to be useful, it must accurately reflect<br />

the status of known at-risk species. Results of scoring<br />

and ranking for 20 selected species using the system can<br />

be found in Tables 4 and 5. These scores are based primarily<br />

on species I have some familiarity with, and in<br />

some cases the ranking should be considered tentative.<br />

A variety of species were used as examples, ranging<br />

from critically endangered endemics such as Artomecon<br />

humilis and Ranunculus aestivalis, sparse widespread<br />

specialists such as Epipactis gigantea, and common endemics<br />

such as Chrysothamnus stylosa, to widespread<br />

Table 4. Selected species of the American Southwest tentatively scored with the ranking system.<br />

The final at-risk score (ARS) is listed. Scores in parentheses reflect uncertainty or lack of data about some aspect of<br />

the scoring. The ARS can vary from 0 (low at-risk) to 18 (critically endangered).<br />

Species Rarity Biology Trend Threats<br />

31<br />

Climate<br />

Change N ARS<br />

Arctomecon humilis 3 3 3 3 3 3 18<br />

Ranunculus aestivalis 3 2 3 3 3 3 17<br />

Ipomopsis sancti-spiritus 3 2 (3) 2 3 3 (16)<br />

Astragalus ampullarioides 3 3 2 2 3 3 16<br />

Pediocactus bradyi 3 3 2 2 3 2 15<br />

Puccinellia parishii 2 (3) 2 3 3 2 (15)<br />

Actaea arizonica 3 3 2 2 3 2 15<br />

Penstemon albomarginatus 3 2 2 2 2 2 13<br />

Camissonia atwoodii 3 3 1 1 3 1 12<br />

Ostrya knowltonii 2 (3) (3) 1 2 1 (12)<br />

Spiranthes diluvialis 2 2 2 1 3 1 11<br />

Cycladenia humilis 2 2 1 1 3 1 10<br />

Epipactus gigantea 2 3 1 2 2 0 10<br />

Carnegiea gigantea 1 3 1 2 2 0 9<br />

Cirsium rydbergii 2 2 1 1 2 1 9<br />

Salix gooddingii 1 2 1 2 3 0 9<br />

Erigeron maguirei 3 2 1 0 1 1 8<br />

Chrysothamnus stylosa 1 1 1 2 1 0 6<br />

Quercus gambelii 0 3 0 1 1 0 5<br />

Pinus ponderosa 0 2 0 0 0 0 2


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

common species like Pinus ponderosa. The conservation<br />

scores range from 18 for A. humilis, 17 for R. aestivalis,<br />

and 9 for E. gigantea, 6 for C. stylosa, and 2 for<br />

P. ponderosa. If the score is in parentheses, this means<br />

that some aspect of its threat status, biology, distribution,<br />

etc., is unknown. Uncertainty resulting from a lack<br />

of data or understanding of the species status and biology<br />

should be included in the ranking system (W. Fertig,<br />

pers. comm. 2009). When an element cannot be<br />

scored due to a lack of data, that element is automatically<br />

scored a 3, thus tending to elevate the species in<br />

the ranking list.<br />

DISCUSSION AND FUTURE WORK<br />

The first aspect of this approach to ranking that<br />

needs to be understood better is how the at-risk score<br />

(ARS) is distributed across a large number of rare taxa.<br />

Are there distinct breaks in the score that reflect underlying<br />

causes of rarity, or are the scores more likely to<br />

exhibit a continuous distribution? If breaks occur, can<br />

they be related to other ranking systems, such as the<br />

IUCN categories of critically endangered, endangered,<br />

and threatened? The significance of a high score is<br />

probably clear in this system as in others, as those species<br />

with high scores are likely to be already endangered.<br />

But we still need to know what kind of vulnerability<br />

to human-caused threats and climate change intermediate<br />

scores might reflect. Ultimately, the system<br />

may need to be refined and altered as ranking of species<br />

is completed and patterns begin to emerge.<br />

Since the ARS reflects a preliminary triage approach<br />

to “potential” vulnerability to near and mid-term threats<br />

(the next 10-50 years), a first use of the system would<br />

be to seek funding for research focused on medium to<br />

high-ranked species to see if in fact they are endangered<br />

or are likely to become endangered. Some previous<br />

work shows that species that initially were considered<br />

rare and in some cases listed under the Endangered<br />

Table 5. General ranking, rarity type, nativity and geographic range for selected species of the American<br />

Southwest. The geographic ranges are derived from McLaughlin (2007; Figure 3).<br />

Species Rarity ARS Score Nativity Geographic Range<br />

Arctomecon humilis RVIII 18 Endemic Mohave Desert<br />

Ranunculus aestivalis RVIII 17 Endemic Colorado Plateau<br />

Ipomopsis sancti-spiritus RVIII (16) Endemic S. Rocky Mountains<br />

Astragalus ampullarioides RVIII 16 Endemic Mohave Desert<br />

Pediocactus bradyi RVIII 15 Endemic Colorado Plateau<br />

Puccinellia parishii RVI (15) Indigenous Southwestern Region<br />

Actaea arizonica RVIII 15 Paleoendemic Colorado Plateau + Madrean<br />

Penstemon albomarginatus RVIII 13 Endemic Mohave Desert<br />

Camissonia atwoodii RVIII 12 Endemic Colorado Plateau<br />

Ostrya knowltonii RVIII (12) Paleoendemic Colorado Plateau + s NM<br />

Spiranthes diluvialis RVI 11 Indigenous Colorado Plateau + Great Plains<br />

Cycladenia humilis RIV 10 Paleoendemic Northern and Southwestern Regions<br />

Epipactis gigantea RVI 10 Indigenous Widespread North American<br />

Carnegiea gigantea RIII 9 Endemic Sonoran Desert<br />

Cirsium rydbergii RIV 9 Endemic Colorado Plateau<br />

Salix gooddingii RII 9 Indigenous Eastern and Southwestern Regions<br />

Erigeron maguirei RVIII 8 Endemic Colorado Plateau<br />

Chrysothamnus stylosa RIII 6 Endemic Colorado Plateau<br />

Quercus gambelii R1 5 Indigenous Southwestern Region<br />

Pinus ponderosa RI 2 Indigenous Northern and Southwestern Regions<br />

32


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Species Act have after further surveys been found to be<br />

relatively common. A good example of this is the<br />

Maguire Daisy, Erigeron maguirei, which was federally<br />

listed as threatened prior to comprehensive surveys of<br />

potential habitat. It has since been found to be fairly<br />

common and is recommended to be upgraded to G3<br />

status (Clark and Tait 2007). Its ARS score (8) is lower<br />

some other widespread species such as Epipactus gigantea<br />

and Salix goodingii.<br />

There are several future directions that can be taken<br />

with this plant rarity ranking system. First, a workshop<br />

needs to be held where botanists with knowledge of<br />

various rare species get together and test the ranking<br />

system with a species list. Such a workshop is tentatively<br />

planned for 2011 to examine and finalize the<br />

G1G2 list of 700 species in the American Southwest. A<br />

second direction is for individual researchers to use the<br />

ranking system in their own area of expertise and determine<br />

how well it can be applied, how relevant it may be,<br />

and what refinements may be needed. It is hoped that<br />

readers of this paper will be interested in doing this. I<br />

encourage those of you who attempt to use the system to<br />

contact me with comments, criticisms, and suggestions.<br />

Finally, if in the future this ranking system, or some version<br />

of it, is deemed to be useful for conservation planning,<br />

it then needs to be incorporated into future action<br />

plans, regional planning efforts, conservation documents,<br />

and heritage databases.<br />

ACKNOWLEDGEMENTS<br />

I would like to thank Walter Fertig and Steve Caicco<br />

for stimulating discussions on how to rank plant rarity,<br />

and to Walt for reviewing an early draft of the manuscript.<br />

Emily Palmquist, Kyle Christy, and Meredith<br />

Jabis provided help in literature searches and data compilation.<br />

LITERATURE CITED<br />

Aizen, M.A., L. Ashworth and G. Leonardo. 2002.<br />

Reproductive success in fragmented habitats: do compatibility<br />

systems and pollination specialization matter?<br />

Journal of Vegetation Science 13: 885-892.<br />

Bevill, R.I. and S.M. Louda. 1999. Comparisons of<br />

related rare and common species in the study of plant<br />

rarity. Conservation Biology 13: 493-498.<br />

Bond, W.J. 1994. Do mutualisms matter? Assessing<br />

the impact of pollinator and disperser disruption on<br />

plant extinction. Philosophical Transactions of the<br />

Royal <strong>Society</strong> of London, B. 344: 83-90.<br />

Brigham, C.A. and M.W. Schwartz (eds.). 2003.<br />

Population viability in plants. Conservation, management,<br />

and modeling of rare plants. Springer, NewYork.<br />

Buchmann, S.L. and G.P. Nabhan. 1996. The forgotten<br />

pollinators. Island Press, Covello, CA.<br />

Byers, D.L. and T.R. Meagher. 1997. A comparison<br />

of demographic characteristics in a rare and a common<br />

species of Eupatorium. Ecological Applications 7: 519-<br />

530.<br />

California <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 2001. Inventory of<br />

rare and endangered plants of California, 6 th Ed. California<br />

<strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> Special Publication No. 1.<br />

Clark, D.J. and D.A. Tait. 2007. Interagency rare<br />

plant team inventory results – 1998 through 2003. Pp.<br />

32-38 In Barlow-Irick, P., J. Anderson and C. McDonald<br />

(eds.). Southwestern rare and endangered plants.<br />

Proceedings of the 4 th Conference. U.S. Forest Service<br />

Proceedings RMRS-P-48CD. Rocky Mountain research<br />

Station. Fort Collins, CO.<br />

Cole, C.T. 2003. Genetic variation in rare and common<br />

plants. Annual Review of Ecology and Systematics<br />

34: 213-237.<br />

Faber-Langendoen, D., L. Master, J. Nichols, K.<br />

Snow, A. Tomaino, R. Bittman, G. Hammerson, B. Heidel,<br />

L. Ramsay, and B. Young. 2009. NatureServe conservation<br />

status assessments: Methodology for assigning<br />

ranks. NatureServe, Arlington, VA. 42 pp.<br />

Ge, S., K.-Q. Wang, D.-Y. Hong, W.-H. Zhang and<br />

Y.-G. Zu. 1999. Comparisons of genetic diversity in the<br />

endangered Adenophora lobophylla and its widespread<br />

congener, A. potaninii. Conservation Biology 13: 509-<br />

513.<br />

Gitzendanner, M.A. and P.S. Soltis. 2000. Patterns of<br />

genetic variation in rare and widespread plant congeners.<br />

American Journal of Botany 87: 783-792.<br />

Harper, J.L. 1981. The meanings of rarity. Pp. 189-<br />

203 In Synge, H. (ed.). The biological aspects of rare<br />

plant conservation. J. Wiley & Sons, New York.<br />

Harper, K.L. 1979. Some reproductive and life history<br />

characteristics of rare plants and implications of<br />

management. Great Basin Naturalist Memoirs 3: 129-<br />

138.<br />

Krupnick, G.A. and W.J. Kress (eds.). 2005. <strong>Plant</strong><br />

conservation. A natural history approach. University of<br />

Chicago Press, Chicago.<br />

Kunin, W.E. and K.J. Gaston (eds.). 1997. The biology<br />

of rarity: causes and consequences of rare-common<br />

differences. Chapman and Hall, New York.<br />

Kunin, W.E. and A. Schmida. 1997. <strong>Plant</strong> reproductive<br />

traits as a function of local, regional and global<br />

abundance. Conservation Biology 11: 183-192.<br />

Kruckeberg, A.R. and D. Rabinowitz. 1985. Biological<br />

aspects of endemism in higher plants. Annual Review<br />

of Ecology and Systematics 16: 447-479.<br />

Kwak, M.M. and R.M. Bekker. 2006. Ecology of<br />

plant reproduction: extinction risks and restoration perspectives<br />

of rare plant species. Pp. 362-386 In Waser,<br />

N.M. and J. Ollerton (eds.). <strong>Plant</strong>-pollinator interactions:<br />

from specialization to generalization. University<br />

of Chicago Press, Chicago.<br />

33


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Lavergne, S., J.D. Thompson, E. Garnier and M. Debussche.<br />

2004. The biology and ecology of narrow endemic<br />

and widespread plants: a comparative study of<br />

trait variation in 20 congeneric pairs. Oikos 107: 505-<br />

518.<br />

Lesica, P. R. Yurkewycz and E.E. Crone. 2006. Rare<br />

plants are common where you find them. American<br />

Journal of Botany 93: 454-459.<br />

Mace, G.M. 1994. Classifying threatened species:<br />

means and ends. Philosophical Transactions of the<br />

Royal <strong>Society</strong> of London, B. 344: 91-97.<br />

Mace, G.M. and R. Lande. 1991. Assessing extinction<br />

threats: toward a reevaluation of IUCN threatened<br />

species categories. Conservation Biology 5: 148-157.<br />

McLaughlin, S.P. 2007. Tundra to tropics: The floristic<br />

plant geography of North America. Sida Botanical<br />

Miscellany No. 30.<br />

Miles, L., A. Grainger and O. Phillips. 2004. The<br />

impact of global climate change on tropical forest biodiversity<br />

in Amazonia. Global Ecology and Biogeography<br />

13: 553-565.<br />

Purdy, B.G., R.J. Bayer and S.E. Macdonald. 1994.<br />

Genetic variation, breeding system evolution, and conservation<br />

of the narrow sand dune endemic Stellaria<br />

arenicola and the widespread S. longipes (Caryophyllaceae).<br />

American Journal of Botany 81: 904-911. et al.<br />

Rabinowitz, D. 1981. Seven forms of rarity. Pp. 205-<br />

217 In Synge, H. (ed.). The biological aspects of rare<br />

plant conservation. J. Wiley & Sons, N.Y.<br />

Robertson, A.W., D. Kelley, J.J. Ladley, and A.D.<br />

Sparrow. 1999. Effects of pollinator loss on endemic<br />

New Zealand mistletoes (Loranthaceae). Conservation<br />

Biology 13: 499-508.<br />

Sakai, A.K., W.L. Wagner and L.A. Mehrhoff. 2002.<br />

Patterns of endangerment in the Hawaiian flora. Systematic<br />

Biology 51: 276-302.<br />

Schemske, D.W., B.C. Husband, M.H. Ruckelhaus,<br />

C. Goodwillie, I.M. Parker and J.G. Bishop. 1994.<br />

Evaluating approaches to the conservation of rare and<br />

endangered plants. Ecology 75: 584-606.<br />

Scott, J.M., F. Davis, B. Csuti, R. Noss, B.<br />

Butterfield, C. Groves, H. Anderson, S. Caicco, F.<br />

D’Erchia, T.C. Edwards, Jr., J. Ulliman and R.G.<br />

Wright. 1993. Gap analysis: a geographic approach to<br />

protection of biological diversity. Wildlife Monographs<br />

No. 123: 3-41.<br />

Sjöström, A. and C.L. Gross. 2006. Life-history<br />

characters and phylogeny are correlated with extinction<br />

risk in the Australian angiosperms. Journal of Biogeography<br />

33: 271-290.<br />

Sodhi, N.S., L.P. Koh, K.S.-H. Peh, H.T.W. Tan,<br />

R.L. Chazdon, R.T. Corlett, T.M. Lee, R.K. Colwell,<br />

B.W. Brook, C.H. Sekercioglu and C.J.A. Bradshaw.<br />

2008. Correlates of extinction proneness in tropical angiosperms.<br />

Diversity and Distributions 14: 1-10.<br />

Stebbins, G.L. and J. Major. 1965. Endemism and<br />

speciation in the California flora. Ecological Monographs<br />

35: 1-34.<br />

Walck, J.L., J.M. Baskin and C.C. Baskin. 1999.<br />

Relative competitive abilities and growth characteristics<br />

of a narrowly endemic and a geographically widespread<br />

Solidago species (Asteraceae). American Journal of<br />

Botany 86: 820-828.<br />

34


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

The Contribution of Cedar Breaks National Monument<br />

to the Conservation of Vascular <strong>Plant</strong> Diversity in <strong>Utah</strong><br />

Walter Fertig and Douglas N. Reynolds,<br />

Moenave Botanical Consulting, Kanab, UT<br />

Abstract. Like most national parks in <strong>Utah</strong>, Cedar Breaks National Monument was initially established to protect its<br />

spectacular scenery rather than to preserve biological diversity. At less than 2500 ha, the monument is one of the<br />

smallest in the region and has a relatively small vascular flora of 354 documented species. Based on conventional<br />

measures of species richness (alpha diversity), Cedar Breaks might not seem like an important component of the protected<br />

area network of <strong>Utah</strong>. However, nearly 10% of the flora of Cedar Breaks is comprised of local or regional endemics<br />

that are mostly restricted to the Claron Formation or volcanic substrates. Many of these are rare species of<br />

high management interest. Our surveys in 2007-2008 documented nearly 1200 point locations for 17 of the monument’s<br />

rarest plants, including first records for Aster welshii and Jamesia americana var. rosea. The monument is<br />

especially significant in terms of beta diversity or complementarity, as it protects 63 plant species that are not otherwise<br />

found in NPS units in the state. As measured by an averaged Jaccard’s Coefficient of Similarity, Cedar Breaks<br />

National Monument has the second most unique flora among the parks in <strong>Utah</strong>.<br />

The Cedar Breaks Amphitheater is a large bowlshaped<br />

valley carved from orange and white limey sandstone<br />

layers of the Eocene Claron Formation on the west<br />

face of Cedar Mountain, about 18 miles east of Cedar<br />

City in southwestern <strong>Utah</strong> (Figure 1). Local Indian<br />

tribes called the area “the circle of painted cliffs” or the<br />

“place where the rocks are sliding down all the time”.<br />

Early Mormon settlers named it Cedar Breaks for the<br />

abundance of juniper (known locally as ‘cedar’) and the<br />

precipitous badland cliffs or breaks. President Franklin<br />

Roosevelt acknowledged the area’s “spectacular cliffs,<br />

canyons, and features of scenic, scientific, and educational<br />

interest” in designating Cedar Breaks as a national<br />

monument under the Antiquities Act in 1933<br />

(Evenden et al. 2002, Fertig 2009b). Administration of<br />

the monument was transferred from Dixie National Forest<br />

to the National Park Service (NPS), to be managed<br />

to “conserve unimpaired” the area’s natural and cultural<br />

resources and values “for the enjoyment of this and future<br />

generations” (NPS 2000).<br />

Protection of native biological diversity was not one<br />

of the rationales for creating Cedar Breaks National<br />

Monument, though this would eventually become an<br />

important part of NPS’s mandate to conserve natural<br />

resources. The botanical significance of the Cedar<br />

Breaks area was just beginning to be discovered in the<br />

early 1930s. Botanists George Goodman and C. Leo<br />

Hitchcock collected the holotypes of Breaks draba<br />

(Draba subalpina) and Cedar Breaks wild buckwheat<br />

(Eriogonum panguicense var. alpestre) from the rim of<br />

the Cedar Breaks Amphitheater in 1930 and Bassett<br />

Maguire added the holotype of Cedar Breaks daisy<br />

(Erigeron proselyticus, or Erigeron sionis var. trilob-<br />

atus) in 1934 (Fertig 2009b). These species are among a<br />

suite of nearly two dozen Claron formation endemics<br />

restricted to the Cedar Breaks area and the vicinity of<br />

Bryce Canyon in south-central <strong>Utah</strong> (Madsen 2001).<br />

Today Cedar Breaks National Monument is part of a<br />

network of highly protected areas that conserve biological<br />

diversity. This network includes other NPS units<br />

(national parks, monuments, recreation areas, and historic<br />

sites), designated wilderness areas, research natural<br />

areas, BLM-managed national monuments, and private<br />

nature preserves such as those managed by The Nature<br />

Conservancy. In <strong>Utah</strong>, these lands cover nearly 14% of<br />

the state (Prior-Magee et al. 2007). Though extensive,<br />

the <strong>Utah</strong> network does not yet capture a representative<br />

sample of the full array of the state’s biological diversity.<br />

Protection remains biased towards common and<br />

widespread species and vegetation types of low economic<br />

use (Fertig 2010a, Prior-Magee et al. 2007).<br />

The purpose of this paper is to examine the contribution<br />

of Cedar Breaks National Monument to the state’s<br />

preserve network by comparing the monument’s floristic<br />

composition, species richness (alpha diversity), degree<br />

of endemism, and number of rare species with that<br />

of other parklands. We hope to demonstrate that despite<br />

the monument’s small size, low alpha diversity, and<br />

relatively homogeneous vegetation, Cedar Breaks is<br />

significant because of its large number of plant species<br />

that are not protected elsewhere (i.e., the monument has<br />

high complementarity or beta diversity). We also hope<br />

to show how comparing annotated species checklists<br />

can be useful for identifying and prioritizing specific<br />

taxa that are missing or under-represented in the preserve<br />

network.<br />

35


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

METHODS<br />

From 2005-2008 we developed a revised checklist of<br />

the vascular plant flora of Cedar Breaks National Monument<br />

(Fertig 2009b, Fertig et al. 2009c). This entailed<br />

re-examination of over 700 specimens from the Cedar<br />

Breaks herbarium, as well as relevant collections housed<br />

at Brigham Young University, the University of Wyoming,<br />

and the New York Botanical Garden’s Virtual<br />

Herbarium. Additional species records were obtained<br />

through a review of previous park checklists, weed surveys,<br />

and vegetation studies (Buchanan 1992, Dewey<br />

and Andersen 2005, Jean and Palmer 1987, Roberts and<br />

Jean 1989, Springer et al. 2006). Field surveys were also<br />

conducted to corroborate unvouchered reports and in<br />

conjunction with a systematic rare plant inventory<br />

(Fertig and Reynolds 2009). The final checklist was annotated<br />

with information on synonyms, taxonomic problems,<br />

population size, growth form, global distribution,<br />

nativity, flowering period, and habitat.<br />

Eighteen rare local or regionally endemic plant species<br />

from Cedar Breaks National Monument and adjacent<br />

portions of the Ashdown Gorge Wilderness Area of<br />

Dixie National Forest were surveyed in 2007-2008<br />

(Fertig and Reynolds 2009). Populations of target species<br />

were mapped using a Global Positioning System<br />

(GPS) device and data were collected on population<br />

size, associated species, habitat conditions, and potential<br />

threats. Most species were mapped as individual points<br />

representing the centrum of relevé-like plots of approximately<br />

25-30 square meters. Populations of Arizona<br />

willow (Salix arizonica) occurred in sufficiently dense<br />

patches to be mapped as polygons.<br />

We assembled additional species checklists for other<br />

NPS units and the BLM’s Grand Staircase-Escalante<br />

National Monument (Figure 2, see Table 3 for citations).<br />

These checklists and our data for Cedar Breaks<br />

National Monument were compiled into a master statewide<br />

checklist in Microsoft excel format. The checklist<br />

followed the taxonomy and nomenclature of Welsh and<br />

others (2008). Only <strong>Utah</strong>-specific data were entered for<br />

those parks that crossed state lines (Dinosaur and<br />

Hovenweep National Monuments and Glen Canyon National<br />

Recreation Area). Unfortunately, complete species<br />

lists are not available for most wilderness areas,<br />

research natural areas, TNC preserves, and other highly<br />

protected areas, so these were excluded from the analysis.<br />

Simple queries were run in excel to compare overall<br />

species richness between parks. Jaccard’s index of similarity<br />

1 was calculated across pairs of parks to quantify<br />

beta diversity.<br />

1 Jaccard’s Coefficient of Similarity<br />

is calculated by the formula<br />

C/(N 1 + N 2 – C), where C = the<br />

number of taxa shared between two<br />

samples, N 1 = the number of taxa in<br />

sample one, and N 2 = the number of<br />

taxa in sample two.<br />

Figure 1. Cedar Breaks National<br />

Monument and Ashdown Gorge<br />

Wilderness Area, Iron County,<br />

<strong>Utah</strong>. Map courtesy of Zion National<br />

Park Resource Management<br />

& Research GIS.<br />

36


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

RESULTS<br />

Species Richness<br />

Based on our re-examination of herbarium specimens,<br />

review of literature, and new field work, 354 species<br />

and varieties of vascular plants (Table 1) are currently<br />

known from Cedar Breaks National Monument<br />

(Fertig 2009b, Fertig et al. 2009c). At least 74 of these<br />

taxa have been discovered since 2005. Overall, the<br />

monument’s flora has increased by 21% since Roberts<br />

and Jean (1989) reported 277 species for the area. The<br />

Figure 2. National Park Service units and BLMmanaged<br />

national monuments in <strong>Utah</strong>. Park acronyms:<br />

ARCH (Arches National Park), BRCA (Bryce Canyon<br />

National Park), CANY (Canyonlands National Park),<br />

CARE (Capitol Reef National Park), CEBR (Cedar<br />

Breaks National Monument), DINO (Dinosaur National<br />

Monument), GCNRA (Glen Canyon National Recreation<br />

Area), GOSP (Golden Spike National Historic<br />

Site), GSENM (Grand Staircase-Escalante National<br />

Monument), HOVE (Hovenweep National Monument),<br />

NABR (Natural Bridges National Monument), RABR<br />

(Rainbow Bridge National Monument), TICA<br />

(Timpanogos Cave National Monument), ZION (Zion<br />

National Park). Map courtesy of Zion National Park<br />

Resource Management & Research GIS.<br />

flora of Cedar Breaks contains 9.7% of the 3659 native<br />

and naturalized plant species documented in <strong>Utah</strong> by<br />

Welsh and others (2008) and 36% of the 156 reported<br />

families.<br />

<strong>Native</strong> species account for 94.9% of the flora of Cedar<br />

Breaks National Monument (336 species). Nearly<br />

82% of the species range widely across western North<br />

America and are common in <strong>Utah</strong> (Table 1). Eighteen<br />

taxa are categorized as local endemics that occupy a<br />

total area of less than 16,500 square kilometers and are<br />

restricted to the immediate vicinity of Cedar Breaks or<br />

adjacent high plateaus of south-central <strong>Utah</strong> (mostly the<br />

Tushar Range and Paunsaugunt Plateau) (Fertig 2009b).<br />

An additional 20 taxa are found only in the Colorado<br />

Plateau area of southern <strong>Utah</strong>, northeastern Arizona,<br />

northwestern New Mexico, and southwestern Colorado.<br />

Together these 38 local and regional endemics account<br />

for 10.7% of the monument’s flora. Just over 2% of the<br />

flora consists of species that occur sporadically across<br />

<strong>Utah</strong> (sparse taxa) or have populations in Cedar Breaks<br />

that are widely isolated from their main, contiguous<br />

range (disjunct) (Table 1).<br />

Only 18 introduced species (those not historically<br />

native to <strong>Utah</strong> or North America) have become established<br />

in the monument, representing 5.1% of the total<br />

flora (Table 1). The percentage of introduced species at<br />

Cedar Breaks is less than half that reported for the entire<br />

flora of <strong>Utah</strong> (13.5%) (Fertig 2007, Welsh et al.<br />

2008). None of the introduced plant taxa in the monument<br />

are listed by the state of <strong>Utah</strong> as official noxious<br />

weeds.<br />

Rare Species<br />

The Conservation Data Center (CDC) of the <strong>Utah</strong><br />

Division of Wildlife Resources (1998) recognizes 22<br />

species from Cedar Breaks National Monument as species<br />

of concern (Table 2). Most of these are local or<br />

regional endemics restricted to the Claron Formation or<br />

species that are widespread outside of <strong>Utah</strong> but have 10<br />

or fewer extant populations in the state. We consider<br />

two additional species from the monument to be deserving<br />

of recognition by the CDC. Madsen’s daisy<br />

(Erigeron vagus var. madsenii) is a southern <strong>Utah</strong> endemic<br />

that was only described as a new taxon in 2008<br />

(Welsh et al. 2008). Rosy cliff jamesia (Jamesia americana<br />

var. rosea) was not recognized as occurring in the<br />

state of <strong>Utah</strong> until we verified populations in Cedar<br />

Breaks and the Ashdown Gorge Wilderness Area in<br />

2008. Previously, populations of this taxon were<br />

thought to represent var. zionis, a local endemic of<br />

southern <strong>Utah</strong> listed by the CDC as a species of concern<br />

and by the US Forest Service and BLM as sensitive.<br />

Var. rosea was formerly known only from California<br />

and Nevada (Fertig and Reynolds 2009, Holmgren and<br />

Holmgren 1989). In all, seven species from Cedar<br />

37


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. Statistical Summary of the Flora<br />

of Cedar Breaks National Monument*<br />

Category<br />

Taxonomic Diversity<br />

Species &<br />

Varieties<br />

Full Species<br />

Only<br />

No. of Taxa<br />

Confirmed<br />

Present<br />

No. of Taxa<br />

Reported<br />

(not confirmed)<br />

Total<br />

347 7 354<br />

335 5 340<br />

Families 56 0 56<br />

Biogeographic Diversity<br />

Introduced 17 1 18<br />

<strong>Native</strong> 330 6 336<br />

Locally<br />

Endemic<br />

Regionally<br />

Endemic<br />

18 0 18<br />

20 0 20<br />

Disjunct 2 0 2<br />

Peripheral 0 0 0<br />

Sparse 6 0 6<br />

Widespread 284 6 290<br />

The number of taxa and plant families is based on Welsh and<br />

others (2008). Biogeographic diversity categories refer to the<br />

distribution of a species within <strong>Utah</strong> and the state’s contribution<br />

to its overall global range (Fertig 2009b). Introduced taxa<br />

are not native to <strong>Utah</strong> or North America but have become<br />

naturalized (breeding on their own without human assistance).<br />

Local Endemics have their entire global range restricted<br />

to an area of less than 16,500 square km (ca 6370 sq<br />

miles, or 1 degree of latitude x 2 degrees of longitude). Regional<br />

Endemics have global ranges of 16,500-250,000<br />

square km (an area about the size of Wyoming). Disjuncts are<br />

isolated from the contiguous portion of their range by a gap<br />

of more than 800 km (ca 500 miles). Peripherals are widespread<br />

globally but occur at the margin of their contiguous<br />

range in <strong>Utah</strong> and occupy less than 5% of the state’s area<br />

(usually only within a few miles of the state border). Sparse<br />

taxa occur widely across <strong>Utah</strong> or North America but their<br />

range within <strong>Utah</strong> is small and patchy, with populations restricted<br />

to specialized or uncommon habitats. Widespread<br />

taxa have global ranges exceeding 250,000 square km and<br />

occur over at least 10% of the state.<br />

* See Addendum for additional species documented since<br />

2009.<br />

Breaks National Monument are presently listed as sensitive<br />

by the US Forest Service or BLM and 14 were once<br />

candidates for listing as Threatened or Endangered under<br />

the US Endangered Species Act (Table 2).<br />

During 2007-2008 we targeted 16 of Cedar Breaks’<br />

rarest local endemics and CDC species of concern<br />

(Table 2) for survey. The number of target species increased<br />

to 18 with the discovery of extant populations of<br />

Madsen’s daisy and the first records of Welsh’s aster<br />

(Aster welshii) for the monument. We recorded at least<br />

one population of 16 of the target species at 546 different<br />

sampling points within Cedar Breaks National<br />

Monument or the Ashdown Gorge Wilderness Area.<br />

Since more than one target species was often present at<br />

each location, we actually documented 1181 different<br />

sample points for these species. For the clonal species<br />

Salix arizonica we delineated 16 discrete polygons in<br />

three main population clusters that cover a total area of<br />

just over one hectare (Fertig and Reynolds 2009).<br />

Ten of our target species occurred in over 10% of our<br />

samples. These species were found mostly on the red or<br />

white limey-sandstone layers of the Claron Formation<br />

along the rim and slopes of the Cedar Breaks Amphitheater.<br />

Cedar Breaks wild buckwheat (Eriogonum panguicense<br />

var. alpestre) was the most widespread and<br />

abundant of the rare species, being found in 36% of all<br />

samples and having a population estimated at 35,200-<br />

100,000 individuals (Fertig and Reynolds 2009). This<br />

plant also has the smallest geographic range of any<br />

taxon in our study, being known only from the Cedar<br />

Breaks area within the monument and the adjacent<br />

Dixie National Forest and Ashdown Gorge Wilderness.<br />

Only three other species were estimated to have populations<br />

of over 10,000 plants: Least lomatium (Lomatium<br />

minimum), Markagunt aster (Aster wasatchensis var.<br />

wasatchensis), and Least spring-parsley (Cymopterus<br />

minimus). The least abundant and most restricted species<br />

in the study area were Rosy cliff jamesia (known<br />

from only about 100 plants in two main areas; Madsen’s<br />

daisy (approximately 400 plants in three main areas),<br />

Reveal’s paintbrush (Castilleja parvula var. revealii<br />

with 500 plants in two main populations), Podunk<br />

groundsel (Senecio malmstenii with about 1500 individuals<br />

in five sites), and Welsh’s aster (with an estimated<br />

1700 plants scattered along Ashdown and Rattle<br />

creeks in the bottom of the amphitheater).<br />

We were unsuccessful in relocating just one of the 18<br />

target species, the Zion draba (Draba asprella var. zionensis).<br />

This species is known from a single herbarium<br />

specimen (Dickman s.n. CEBR) collected from “Cedar<br />

Breaks National Monument” in 1977. Unfortunately,<br />

nothing more precise is known about the original collection<br />

site. Zion draba occurs commonly in Zion National<br />

Park on Navajo Sandstone cliffs and canyons. Comparable<br />

Navajo Sandstone outcrops are not exposed at Cedar<br />

38


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 2. Vascular <strong>Plant</strong> Species of Concern of Cedar Breaks National Monument<br />

Family<br />

Apiaceae<br />

(Umbelliferae)<br />

Apiaceae<br />

(Umbelliferae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Asteraceae<br />

(Compositae)<br />

Brassicaceae<br />

(Cruciferae)<br />

Species/<br />

Common Name<br />

Cymopterus minimus<br />

Least spring-parsley<br />

Lomatium minimum<br />

Least lomatium<br />

Agoseris glauca var. agrestis*<br />

Field agoseris<br />

Antennaria pulcherrima*<br />

Showy pussytoes<br />

Aster wasatchensis var.<br />

wasatchensis<br />

Markagunt aster<br />

Aster welshii<br />

Welsh’s aster<br />

Erigeron sionis var. trilobatus<br />

Cedar Breaks daisy<br />

Erigeron vagus var. madsenii<br />

Madsen’s daisy<br />

Haplopappus zionis<br />

Cedar Breaks goldenbush<br />

Machaeranthera commixta*<br />

Bigelow’s aster<br />

Senecio malmstenii<br />

Podunk groundsel<br />

Townsendia montana var.<br />

minima<br />

Bryce Canyon townsendia<br />

Draba asprella var. zionensis<br />

Zion draba<br />

TNC Rank Legal & UTCDC<br />

Status<br />

G1G2Q/<br />

S1S2<br />

G3/S3<br />

USFS: Sensitive<br />

USFWS: former C2<br />

UTCDC: Rare<br />

USFWS: former 3C<br />

UTCDC: Watch<br />

G5T5/S1S2 UTCDC: Taxonomic<br />

Problems<br />

G5?/S1<br />

Abundance in CEBR<br />

Ca 12,500 plants, in<br />

35% of sample plots<br />

Ca 33,600 plants, in<br />

13.2% of sample plots<br />

Not known<br />

UTCDC: Peripheral Not known<br />

G2/S2 UTCDC: Watch Ca 15,100 plants in<br />

17.8% of sample plots<br />

G2/S2 UTCDC: Watch Ca 1700 plants in 4.6%<br />

of sample plots<br />

G2/S2<br />

USFWS: former C2<br />

UTCDC: Watch<br />

Ca 4200 plants in<br />

13.6% of sample plots<br />

G4T?/SNR Ca 4000 plants in 2%<br />

of sample plots<br />

G2/S2<br />

G4G5T3?/<br />

S3?<br />

G1/S1?<br />

G4T3/S3<br />

UT BLM: Sensitive<br />

USFWS: former C2<br />

UTCDC: Watch<br />

UTCDC: Watch<br />

USFS: Sensitive<br />

USFWS: former C2<br />

UTCDC: Rare<br />

USFWS: former C2<br />

UTCDC: Watch<br />

G4T3?/S3? USFWS: former 3C<br />

UTCDC: Watch<br />

Ca 4100 plants in<br />

14.3% of sample plots<br />

Not known<br />

Ca 1500 plants in 4%<br />

of sample plots<br />

Ca 4100 plants in 11%<br />

of sample plots<br />

Not known, may be<br />

falsely reported<br />

Brassicaceae<br />

(Cruciferae)<br />

Draba subalpina<br />

Breaks draba<br />

G3/S3<br />

USFWS: former 3C<br />

UTCDC: Watch<br />

Ca 8700 plants in 22%<br />

of sample plots<br />

Brassicaceae<br />

(Cruciferae)<br />

Caryophyllaceae<br />

Equisetaceae<br />

Physaria rubicundula var.<br />

rubicundula*<br />

Breaks bladderpod<br />

Silene petersonii<br />

Peterson’s campion<br />

Equisetum variegatum*<br />

Northern scouring-rush<br />

G3//S3<br />

USFWS: former 3C<br />

UTCDC: Watch<br />

G2G3/S2S3 USFS: Sensitive;<br />

USFWS: former C2<br />

UTCDC: Watch<br />

G5/S1<br />

Not known<br />

Ca 2900 plants in<br />

12.1% of plots<br />

UTCDC: Peripheral Not known<br />

39


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 2. continued<br />

Family<br />

Species/<br />

Common Name<br />

TNC Rank Legal & UTCDC<br />

Status<br />

Abundance in CEBR<br />

Fabaceae<br />

(Leguminosae)<br />

Hydrangeaceae<br />

(Saxifragaceae)<br />

Polemoniaceae<br />

Polygonaceae<br />

Pyrolaceae<br />

(Ericaceae)<br />

Salicaceae<br />

Scrophulariaceae<br />

(Orobanchaceae)<br />

Astragalus limnocharis var.<br />

limnocharis<br />

Navajo Lake milkvetch<br />

Jamesia americana var. rosea<br />

Rosy jamesia<br />

Ipomopsis tridactyla<br />

Tushar gilia<br />

Eriogonum panguicense var.<br />

alpestre<br />

Cedar Breaks wild buckwheat<br />

Pyrola picta*<br />

White-veined wintergreen<br />

Salix arizonica<br />

Arizona willow<br />

Castilleja parvula var. revealii<br />

Reveal’s paintbrush<br />

G2T1/S1<br />

USFS: Sensitive<br />

USFWS: former C2<br />

Ca 4100 plants in<br />

15.9% of sample plots<br />

G5T3/SNR Ca 100 plants in 1.1%<br />

of sample plots<br />

G5T2/S2 UTCDC: Rare Ca 2800 plants in 9.3%<br />

of sample plots<br />

G3T2T3Q/<br />

SSYN<br />

G4G5/S1<br />

G2G3/S2<br />

G2/S2<br />

USFWS: former 3C<br />

UTCDC: Taxonomic<br />

Problems<br />

35,200-100,000 plants<br />

estimated, in 36% of<br />

sample plots<br />

UTCDC: Infrequent Not known<br />

USFS: Sensitive<br />

USFWS: former<br />

Candidate<br />

UTCDC: Rare<br />

USFS: Sensitive<br />

USFWS: former C2<br />

16 small to medium<br />

sized clones in 3 main<br />

population clusters<br />

covering 1.06 ha<br />

Ca 500 plants, in 4.8%<br />

of sample plots<br />

*Species not surveyed in 2007-2008.<br />

Derived from <strong>Utah</strong> Department of Wildlife Resources (1998), Fertig (2009b), Fertig and Reynolds (2009), and Fertig and others<br />

(2009c). Codes: TNC rank assesses abundance and conservation priority on a scale of 1-5 (1 being extremely vulnerable and 5<br />

being secure) for full species (G) and varieties or subspecies (T) across their entire range and within each state (S). A “?” indicates<br />

uncertainty in the rank, Q = taxonomic questions, NR = not ranked, and SYN = species is considered a synonym and not<br />

ranked under the given name. Under legal status, USFWS = US Fish and Wildlife Service. C2 = category 2 candidate (a former<br />

category used for taxa that might warrant being proposed for Threatened or Endangered status following additional research). 3C<br />

= category 3 candidates (species dropped from consideration for listing). USFS = US Forest Service. BLM = Bureau of Land<br />

Management. UTCDC status includes conservation categories adopted by the state natural heritage program to prioritize endemic<br />

and rare plant taxa (<strong>Utah</strong> Division of Wildlife Resources 1998). These categories include: Rare (plants with rangewide viability<br />

concerns), Watch (regional endemics without rangewide viability concerns), Peripheral (rare or uncommon in <strong>Utah</strong>, but more<br />

common rangewide), Infrequent (plants occur infrequently over western US), and Taxonomic Problems (validity of species, subspecies,<br />

or variety has been questioned).<br />

Breaks National Monument, though similar sandstone<br />

cliffs of the Straight Cliffs or Wahweap formations are<br />

present in the bottom of Ashdown Canyon. These areas<br />

were searched in 2008, but no populations were found.<br />

It is possible that the label of the specimen is erroneous<br />

and the collection was actually made in Zion National<br />

Park (Fertig and Reynolds 2009).<br />

Similarity to Other Park Floras<br />

Cedar Breaks National Monument ranks tenth out of the<br />

14 NPS and BLM managed parklands considered in this<br />

study in both area and species richness (Table 3). As of<br />

2008, Grand Staircase-Escalante National Monument is<br />

the largest protected area at over 761,000 ha and has the<br />

highest number of vascular plant species with 999 2 . In<br />

general, vascular plant species richness is positively correlated<br />

with total area, with the main exception being<br />

Zion National Park with the second highest number of<br />

species (991) but ranking sixth in total area. When park<br />

2 More than 10 new species were documented in Zion National<br />

Park in 2009, allowing it to pass Grand Staircase-Escalante<br />

National Monument as the <strong>Utah</strong> parkland with the highest<br />

vascular plant species richness. See Fertig and others (<strong>2012</strong>)<br />

for updates on the flora of each NPS unit assessed in this<br />

study.<br />

40


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 3. <strong>Number</strong> of Vascular <strong>Plant</strong> Taxa in National Park Service and Bureau of Land<br />

Management Parks, Monuments, Historic Sites, and Recreation Areas in <strong>Utah</strong><br />

Management Area Size (ha) Total #<br />

of Taxa<br />

# of Rare<br />

Taxa a<br />

# of<br />

Endemic<br />

Taxa<br />

# of<br />

Unique<br />

Taxa b<br />

# of Taxa<br />

per log<br />

(area)<br />

Source<br />

Arches NP 30,966 524 32 82 12 50.7 Fertig et al. 2009a,<br />

2009c<br />

Bryce Canyon NP 14,502 587 51 71 28 61.3 Fertig and Topp<br />

2009<br />

Canyonlands NP 136,610 600 55 107 16 50.8 Fertig et al. 2009b,<br />

2009c<br />

Capitol Reef NP 97,895 888 62 142 90 77.3 Fertig 2009a<br />

Cedar Breaks NM 2,491 354 24 38 63 45.3 Fertig 2009b, Fertig<br />

et al. 2009c<br />

Dinosaur NM 85,096 757 (485<br />

in UT)<br />

Glen Canyon NRA 505,868 889 (863<br />

in UT)<br />

80 76 69 (UT) 66.7 Fertig 2009c, Fertig<br />

et al. 2009c<br />

60 176 73 (UT) 67.7 Hill 2005, Spence<br />

2005<br />

Golden Spike NHS 1,107 149 6 6 20 21.3 Fertig 2009d, Fertig<br />

et al. 2009c<br />

Grand Staircase-<br />

Escalante NM<br />

761,070 999 142 193 68 73.8 Fertig 2005 & unpublished<br />

data<br />

Hovenweep NM 318 340 (240<br />

in UT) c 17 31 6 59.0 Fertig 2009e<br />

Natural Bridges NM 3,009 428 21 72 6 53.4 Fertig 2009f<br />

Rainbow Bridge NM 65 224 12 33 4 53.7 Fertig 2010b<br />

Timpanogos Cave NM 101 235 8 15 39 50.9 Fertig and Atwood<br />

2009<br />

Zion NM 59,900 991 189 133 199 90.1 Fertig and Alexander<br />

2009<br />

a Derived from <strong>Utah</strong> Division of Wildlife Resources (1998)<br />

B<br />

Defined as species found in only one of the 14 parklands considered here<br />

c Erroneously reported as 214 taxa in <strong>Utah</strong> in Fertig (2009e)<br />

41


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

size is normalized by taking the natural log of area,<br />

Zion National Park emerges as the largest flora in the<br />

study area at 90.1 species/ln(area), followed by Capitol<br />

Reef National Park and Grand Staircase-Escalante<br />

National Monument (Table 3). Cedar Breaks National<br />

Monument drops to thirteenth in alpha diversity at<br />

45.3 species/ln(area) when area is normalized, exceeding<br />

only Golden Spike National Historic Site.<br />

In our sample, Cedar Breaks National Monument<br />

shares the most vascular plant taxa in common with<br />

Bryce Canyon National Park (227 species), Zion National<br />

Park (202 species), and Grand Staircase-<br />

Escalante National Monument (172 species) (Table 4).<br />

Not surprisingly, these are the three parklands closest<br />

in proximity to Cedar Breaks (Figure 2). The monument<br />

shares the fewest species in common with<br />

Golden Spike National Historic Site (29 species),<br />

Rainbow Bridge National Monument (32 species), and<br />

the <strong>Utah</strong> sections of Hovenweep National Monument<br />

(37 species). Factoring in discrepancies in the relative<br />

sizes of different park floras using Jaccard’s Coefficient<br />

of Similarity (JCS), Cedar Breaks National<br />

Monument is still most similar to Bryce Canyon National<br />

Park (JCS = 0.318), but is significantly less<br />

similar to Zion National Park and Grand Staircase-<br />

Escalante National Monument (JCS = 0.177 and<br />

0.146, respectively). Timpanogos Cave National<br />

Monument, which shared only 75 species in common<br />

with Cedar Breaks, has a Jaccard’s coefficient equal to<br />

that of Grand Staircase, despite being much farther<br />

distant. Based on the Jaccard’s Coefficient, Cedar<br />

Breaks National Monument is least similar to Rainbow<br />

Bridge National Monument and Golden Spike<br />

National Historic Site (Table 4). Cedar Break’s average<br />

coefficient of similarity among the other thirteen<br />

parklands in the study is 0.128. This is the second<br />

lowest value among all the parks analyzed, and is<br />

bested only by Golden Spike National Historic Site<br />

(average JCS = 0.121). By comparison, Canyonlands<br />

National Park (average JCS = 0.326) and Grand Staircase-Escalante<br />

National Monument (average JCS =<br />

0.312) share the most species on average with other<br />

<strong>Utah</strong> parklands.<br />

The low Jaccard’s Coefficient value for Cedar<br />

Breaks is a consequence of the area’s low overall species<br />

richness and the relatively high number of plant<br />

taxa that are unique to the monument. Of the 354 taxa<br />

documented for Cedar Breaks National Monument, 63<br />

species (17.8% of the local flora) are not found in any<br />

of the other protected areas in this study (Table 3).<br />

Only Zion National Park has a higher percentage<br />

(19.9%) of its flora that is protected nowhere else in<br />

the state. Ten of the 24 rare species of Cedar Breaks<br />

are protected only in the monument and another ten<br />

occur just in Cedar Breaks and Bryce Canyon National<br />

Park.<br />

DISCUSSION<br />

Creating a protected area network that contains full<br />

representation of the entire array of native biological<br />

diversity is one of the core principles of contemporary<br />

conservation biology (Groves et al. 2002, Margules and<br />

Pressey 2000, Margules and Sarkar 2007, Scott et al.<br />

1993, Stein et al. 2000). The foundation of such a network<br />

already exists in <strong>Utah</strong>, consisting of national<br />

parks, monuments, recreation areas, historic sites, designated<br />

wilderness, research natural areas, national<br />

wildlife refuges, and Nature Conservancy preserves.<br />

Unfortunately, many of the best protected sites in the<br />

state were originally created for their scenic and historic<br />

values or recreation potential rather than the preservation<br />

of biological diversity. Significant gaps remain in<br />

the state’s protected area network, especially for many<br />

rare and endemic species and plants from lowland habitats<br />

that have been largely converted to agriculture or<br />

residential use (Fertig 2010a, Prior-Magee et al. 2007).<br />

Building a modern preserve network is expensive<br />

and difficult, requiring costly land acquisition, changes<br />

in economic or regulatory policy, and expenditure of<br />

political capital. Because of these costs, it is critical that<br />

conservation strategies be both effective and efficient<br />

(Margules and Sarkar 2007, Stein et al. 2000). To maximize<br />

efficiency, conservation biologists often focus<br />

their efforts on preserving hotspots of unusually high<br />

species richness, representative examples of major<br />

vegetation types (that serve as surrogates for all biodiversity),<br />

and areas with high complementarity. Hotspots<br />

and vegetation types are especially useful for capturing<br />

broad swaths of diversity when building networks from<br />

scratch, but can become less efficient as preserve systems<br />

grow and gaps become more obvious (Williams et<br />

al. 1996). With its emphasis on species that tend to be<br />

rare and localized, complementarity can be an effective<br />

alternative to hotspots and vegetation types when filling<br />

specific holes in the network (Margules and Sarkar<br />

2007, Williams et a. 1996).<br />

Cedar Breaks National Monument has been part of<br />

the <strong>Utah</strong> protected area network for over 75 years. But<br />

if Cedar Breaks were not already protected, would it be<br />

a worthy addition to the network? In terms of overall<br />

species richness, the answer at first glance appears to be<br />

no. The monument contains only 354 vascular plant<br />

species, compared to the average of 558 taxa for the<br />

other parklands considered in this study. The area’s low<br />

alpha diversity can be attributed to the monument’s<br />

small size and relatively homogeneous vegetation<br />

(especially compared to larger parks in the Colorado<br />

Plateau portion of the state). Relative to other parks,<br />

42


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Cedar Breaks also has a fairly low number of endemic<br />

and rare species.<br />

It is in terms of complementarity that Cedar Breaks<br />

makes its primary contribution to the state’s protected<br />

area network. Jaccard’s Coefficient of Similarity is a<br />

useful formula for measuring complementarity as it<br />

weights similarities in floristic composition between<br />

areas by discrepancies in total species richness. With an<br />

average JCS of 0.128, Cedar Breaks National Monument<br />

has the most dissimilar flora of any parkland considered<br />

in this study other than Golden Spike National<br />

Historic Site. The monument’s low JCS score is a consequence<br />

of its relatively high number of unique plant<br />

species (63 taxa or nearly 18% of the local flora) that<br />

are not protected elsewhere in the state. While many of<br />

these unique species are Claron endemics, the majority<br />

are taxa restricted to elevations above 3000 meters. The<br />

average elevation of Cedar Breaks is 2800 meters, a figure<br />

that exceeds the highest elevation of all other parks<br />

in the <strong>Utah</strong> reserve network. The 14 parklands in the<br />

preserve network analyzed here protect 2007 of <strong>Utah</strong>’s<br />

3659 native and naturalized vascular plant species<br />

(54.8% of the total state flora). Cedar Breaks’ unique<br />

contributions account for 3.1% of the total. Ten of the<br />

24 rare plant species recognized for Cedar Breaks and 9<br />

of the 38 local or regional endemics are among the species<br />

protected nowhere else.<br />

The flora of Cedar Breaks is most similar to that of<br />

Bryce Canyon National Park, with the two parks sharing<br />

227 plant taxa. This degree of similarity is not surprising<br />

given the proximity of the areas and their similar<br />

elevation, vegetation, and geology (both parks have extensive<br />

outcrops of reddish-orange Claron Formation<br />

badlands). Although there is redundancy in the protection<br />

of some species in both parks, this is not necessarily<br />

disadvantageous, as having multiple populations in the<br />

protected area network can reduce the risk of catastrophic<br />

loss and enhance the preservation of multiple genotypes<br />

(Noss and Cooperrider 1994, Margules and Sarkar<br />

2007). Similar redundancy occurs among common montane<br />

forest and meadow species found in both Cedar<br />

Breaks and Timpanogos Cave national monuments,<br />

which despite their great distance from each other share<br />

similar vegetation types and a relatively high Jaccard’s<br />

Coefficient of Similarity.<br />

Based on our analysis of vascular plant checklists<br />

from selected parklands, nearly 45% of the native and<br />

naturalized flora of <strong>Utah</strong> is not represented in the state’s<br />

existing preserve network. Some of the omissions may<br />

be an artifact of incomplete sampling or data synthesis,<br />

especially of wilderness areas, research natural areas,<br />

national wildlife refuges, private nature preserves, and<br />

other protected areas that could not be included in this<br />

analysis for lack of data. Other holes will only be filled<br />

by targeting specific plant taxa identified by analyzing<br />

system-wide complementarity. Fortunately, many of the<br />

missing species occur in specific geographic areas or<br />

habitat types which are, themselves, poorly represented<br />

in the current network. Fertig (2010a) identified just 12<br />

geographic areas that, if protected, would capture 70%<br />

of the missing plant species in the preserve network on<br />

<strong>Utah</strong>’s Colorado Plateau. These are mostly “hotspots”<br />

of unprotected endemism and include the La Sal, Abajo,<br />

Henry, Tushar, Boulder, and Pine Valley mountains,<br />

Book Cliffs, Tavaputs, and Fish Lake plateaus, Uinta<br />

Basin, Sevier Valley, and San Rafael Swell. Statewide,<br />

important gaps also exist in the Great Basin, lowland<br />

riparian areas, and the foothills and montane zones of<br />

northern mountains. Future additions to the network<br />

may well be sites like Cedar Breaks that are of relatively<br />

small size or modest species richness but which have<br />

significant beta diversity.<br />

For any reserve network to function, it will be increasingly<br />

important to keep score of what species are<br />

represented, how many populations are captured, and<br />

whether these populations are of sufficient size or quality<br />

to persist. Annotated species lists and databases of<br />

distributions are critical tools for identifying gaps in the<br />

reserve network. Efficient planning and implementation<br />

of a complete reserve system requires that intelligent<br />

choices be made in selecting geographic areas and species<br />

to target for inclusion. Measuring complementarity,<br />

as we have done for Cedar Breaks and other parklands,<br />

is a key step in the prioritization process.<br />

ACKNOWLEDGEMENTS<br />

We would like to thank Paul Roelandt, Superintendent<br />

of Cedar Breaks National Monument, for encouragement<br />

and support; Cynthia Wanschura and Matt Betenson,<br />

formerly of the Zion National Park GIS lab, for<br />

help with maps; and Amy Tendick and Sarah Topp of<br />

the NPS Northern Colorado Plateau Network for sharing<br />

data on new species occurrences. Thanks also to our 2<br />

anonymous reviewers for helpful suggestions.<br />

REFERENCES<br />

Buchanan, H. 1992. Wildflowers of southwestern<br />

<strong>Utah</strong>. Bryce Canyon Natural History Association and<br />

Falcon Press, Helena, MT. 119 pp.<br />

Dewey, S. and K. Andersen. 2005. An inventory of<br />

invasive non-native plants in Cedar Breaks National<br />

Monument (2004): Final report. <strong>Utah</strong> State University.<br />

Weed Science Research Project #SD0515A.<br />

Evenden, A., M. Miller, M. Beer, E. Nance, S. Daw,<br />

A. Wight, M. Estenson, and L. Cudlip. 2002. Northern<br />

Colorado Plateau Vital Signs Network and Prototype<br />

Cluster, Plan for Natural Resources Monitoring: Phase 1<br />

Report, October 1, 2002 [two volumes]. National Park<br />

Service, Northern Colorado Plateau Network, Moab,<br />

UT. 138 pp. + app.<br />

43


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Fertig, W. 2005. Annotated checklist of the flora of<br />

Grand Staircase-Escalante National Monument. Moenave<br />

Botanical Consulting, Kanab, UT. 54 pp.<br />

Fertig, W. 2007. Introduced and naturalized plants of<br />

<strong>Utah</strong>. Sego Lily 30(5):7-11.<br />

Fertig, W. 2009a. Annotated checklist of vascular<br />

flora: Capitol Reef National Park. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/154. 172 pp.<br />

Fertig, W. 2009b. Annotated checklist of vascular<br />

flora: Cedar Breaks National Monument. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/173.<br />

92 pp.<br />

Fertig, W. 2009c. Annotated checklist of vascular<br />

flora: Dinosaur National Monument. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/225. 182 pp.<br />

Fertig, W. 2009d. Annotated checklist of vascular<br />

flora: Golden Spike National Historic Site. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/206.<br />

50 pp.<br />

Fertig, W. 2009e. Annotated checklist of vascular<br />

flora: Hovenweep National Monument. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/207.<br />

112 pp.<br />

Fertig, W. 2009f. Annotated checklist of vascular<br />

flora: Natural Bridges National Monument. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/155.<br />

96 pp.<br />

Fertig, W. 2010a. Finding gaps in the protected area<br />

network in the Colorado Plateau: A case study using<br />

vascular plant taxa in <strong>Utah</strong>. In: Van Riper III, C., B.F.<br />

Wakeling, and T.D. Sisk, eds. The Colorado Plateau IV:<br />

Shaping Conservation Through Science and Management.<br />

University of Arizona Press, Tucson, AZ. Pp<br />

109-119.<br />

Fertig, W. 2010b. The flora of Rainbow Bridge National<br />

Monument. Sego Lily 33(6):4-14.<br />

Fertig, W. and J. Alexander. 2009. Annotated checklist<br />

of vascular flora: Zion National Park. Natural Resource<br />

Technical Report NPS/NCPN/NRTR-2009/157.<br />

196 pp.<br />

Fertig, W. and N.D. Atwood. 2009. Annotated<br />

checklist of vascular flora: Timpanogos Cave National<br />

Monument. Natural Resource Technical Report NPS/<br />

NCPN/NRTR-2009/167. 76 pp.<br />

Fertig, W. and D.N. Reynolds. 2009. Survey of rare<br />

plants of Cedar Breaks National Monument: Final report.<br />

CPCESU Cooperative Agreement #H1200-004-<br />

0002. Colorado Plateau Cooperative Ecosystem Studies<br />

Unit and Southern <strong>Utah</strong> University. 93 pp.<br />

Fertig, W. and S. Topp. 2009. Annotated checklist of<br />

vascular flora: Bryce Canyon National Park. Natural<br />

Resource Technical Report NPS/NCPN/NRTR-<br />

2009/153. 130 pp.<br />

Fertig, W., S. Topp, and M. Moran. 2009a. Annotated<br />

checklist of vascular flora: Arches National Park.<br />

Natural Resource Technical Report NPS/NCPN/NRTR-<br />

2009/220. 117 pp.<br />

Fertig, W., S. Topp, and M. Moran. 2009b. Annotated<br />

checklist of vascular flora: Canyonlands National<br />

Park. Natural Resource Technical Report NPS/NCPN/<br />

NRTR-2009/221. 130 pp.<br />

Fertig, W., S. Topp, M. Moran, and R. Weissinger.<br />

2009c. New vascular plant species discoveries in the<br />

Northern Colorado Plateau Network: 2008 update. Moenave<br />

Botanical Consulting, Kanab, UT. 24 pp.<br />

Fertig, W., S. Topp, M. Moran, T. Hildebrand, J. Ott,<br />

and D. Zobell. <strong>2012</strong>. Vascular plant species discoveries<br />

in the Northern Colorado Plateau Network: Update<br />

for 2008-2011. Natural Resources Technical Report<br />

NPS/NCPN/NRTR –<strong>2012</strong>/582.<br />

Groves, C.R., D.B. Jensen, L.L. Valutis, K.H. Redford,<br />

M.L. Shaffer, J.M. Scott, J.V. Baumgartner, J.V.<br />

Higgins, M.W. Beck, and M.G. Anderson. 2002. Planning<br />

for biodiversity conservation: Putting conservation<br />

science into practice. Bioscience 52(6):499-512.<br />

Hill, M. 2005. Vascular flora of Glen Canyon National<br />

Recreation Area, <strong>Utah</strong> and Arizona. Master’s<br />

thesis, Northern Arizona University, Flagstaff, AZ.<br />

Holmgren, N.H. and P.K. Holmgren. 1989. A taxonomic<br />

study of Jamesia (Hydrangeaceae). Brittonia 41<br />

(4):335-350.<br />

Jean, C. and B. Palmer. 1987. Vascular plant species<br />

list, Cedar Breaks National Monument. 8 pp.<br />

Madsen, M. 2001. A floristic study of the Tertiary<br />

Claron Formation in the Paunsaugunt region of southern<br />

<strong>Utah</strong>. Master’s thesis, Department of Botany and Range<br />

Science, Brigham Young University, Provo, UT.<br />

Margules, C.R. and R.L. Pressey. 2000. Systematic<br />

conservation planning. Nature 405:243-253.<br />

Margules, C.R. and S. Sarkar. 2007. Systematic Conservation<br />

Planning. Cambridge University Press, Cambridge,<br />

UK. 270 pp.<br />

National Park Service (NPS). 2000. Management<br />

policies. US Department of Interior, National Park Service,<br />

Washington, DC. Publication # NPS DI416. 137<br />

pp.<br />

Noss, R.F. and A.Y. Cooperrider. 1994. Saving Nature’s<br />

Legacy: Protecting and Restoring Biodiversity.<br />

Island Press, Washington, DC. 416 pp.<br />

Prior-Magee, J.S., K.G. Boykin, D.F. Bradford, W.G.<br />

Kepner. J.H. Lowry, D.L. Schrupp, K.A. Thomas, and<br />

B.C. Thompson, editors. 2007. Southwest Regional Gap<br />

Analysis Project Final Report. US Geological Survey,<br />

Gap Analysis Program, Moscow, ID.<br />

Roberts, D.W. and C. Jean. 1989. <strong>Plant</strong> community<br />

and rare and exotic species distribution and dynamics in<br />

Cedar Breaks National Monument. Department of Forest<br />

Resources and Ecology Center, <strong>Utah</strong> State University,<br />

Logan, UT. 144 pp.<br />

44


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Scott, J.M., F. Davis, B. Csuti, R. Noss, B.<br />

Butterfield, C. Groves, H. Anderson, S. Caicco, F.<br />

D’Erchia, T.C. Edwards, J. Ulliman, and G. Wright.<br />

1993. Gap analysis: a geographic approach to protection<br />

of biological diversity. Wildlife Monographs 123. 41<br />

pp.<br />

Spence, J.R. 2005. Notes on significant collections<br />

and additions to the flora of Glen Canyon National Recreation<br />

Area, <strong>Utah</strong> and Arizona, between 1992 and<br />

2004. Western North American Naturalist 65(1):103-<br />

111.<br />

Springer, A.E., L.E. Stevens, and R. Harms. 2006.<br />

Inventory and classification of selected National Park<br />

Service springs on the Colorado Plateau. Northern Arizona<br />

University, Flagstaff, AZ. NPS Cooperative<br />

agreement # CA 1200-99-009; Task # NAU-118.<br />

Stein, B.A., L.S. Kutner, and J.S. Adams. 2000. Precious<br />

Heritage, the Status of Biodiversity in the United<br />

States. The Nature Conservancy and Association for<br />

Biodiversity Information, Oxford University Press, New<br />

York. 399 pp.<br />

<strong>Utah</strong> Division of Wildlife Resources. 1998. Inventory<br />

of sensitive species and ecosystems in <strong>Utah</strong>. Endemic<br />

and rare plants of <strong>Utah</strong>: An overview of their distribution<br />

and status. Report prepared for the <strong>Utah</strong> Reclamation<br />

Mitigation and Conservation Commission and<br />

US Department of the Interior. 566 pp. + app.<br />

Welsh, S.L., N.D. Atwood, S. Goodrich, and L.C.<br />

Higgins. 2008. A <strong>Utah</strong> Flora, 2004-2008 summary<br />

monograph, fourth edition, revised. Brigham Young<br />

University Print Services, Provo, UT. 1019 pp.<br />

Williams, P., D. Gibbons, C. Margules, A. Rebelo,<br />

C. Humphries, and R. Pressey. 1996. A comparison of<br />

richness hotspots, rarity hotspots and complementarity<br />

areas for conserving diversity using British birds. Conservation<br />

Biology 10:155-174.<br />

Addendum<br />

Since this paper was written in the summer of 2009, 31 new vascular plant taxa have been collected or reported for<br />

Cedar Breaks National Monument, increasing the flora to 385 species and varieties (Fertig et al. <strong>2012</strong>). Among the<br />

more uncommon or noteworthy additions to the monument's flora are Botrychium lunaria (sparse in <strong>Utah</strong> and the first<br />

species of Ophioglossaceae for Cedar Breaks), Cirsium clavatum var. clavatum (a regional endemic, first collected in<br />

1982 and located in a search of specimens at the Brigham Young University herbarium), and Penstemon caespitosus<br />

var. suffruticosus (a local endemic of southern <strong>Utah</strong>).<br />

In recent years, local political leaders in Iron County, <strong>Utah</strong>, have proposed changing Cedar Breaks from a national<br />

monument to a national park and having the park annex the Ashdown Gorge Wilderness on its western boundary. I<br />

conducted a floristic survey of the Ashdown Gorge Wilderness in the summer of 2009 (Fertig 2009g) and documented<br />

308 vascular plant taxa. Of these species 247 were previously known from Cedar Breaks National Monument,<br />

while 61 were new species for the local area. If the Ashdown Gorge Wilderness Area were added to Cedar<br />

Breaks National Monument the total flora would increase to 426 taxa (Fertig 2009g).<br />

Fertig, W. 2009g. Vascular plant flora of the Ashdown Gorge Wilderness Area and additions to the flora of Cedar<br />

Breaks National Monument. Moenave Botanical Consulting, Kanab, UT. 45 pp.<br />

45


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Studying the Seed Bank Dynamics of Rare <strong>Plant</strong>s<br />

Susan E. Meyer,<br />

USDA Forest Service Shrub Lab, Provo, UT<br />

Abstract. Seed bank studies are important for understanding the population biology of rare plants, especially shortlived<br />

species from unpredictable environments. Seed retrieval studies are used to determine how long seeds persist in<br />

the soil and how seed dormancy changes through time. Even short-term studies (three to five years) can help determine<br />

whether the seed bank is transient or persistent and characterize the shape of the seed depletion trajectory. Laboratory<br />

germination studies can provide clues about seed bank persistence, and in situ seed bank studies can quantify<br />

seed bank size, but only retrieval studies provide the information needed to inform population viability analysis.<br />

In order to assess the status of a rare plant population,<br />

botanists need some way to measure and quantify<br />

demographic (life history) variables. This can be as<br />

straightforward as quantifying the number of plants present<br />

in an area using a simple monitoring protocol, or it<br />

can involve classifying the individuals present by age,<br />

size or stage (for example, seedling, juvenile, vegetative<br />

adult, reproductive adult). It can also include an evaluation<br />

of the potential contribution of an individual to the<br />

next generation, that is, its reproductive output in terms<br />

of numbers of seeds or vegetative ramets produced and<br />

the quality of those seeds or ramets. If the plants are<br />

permanently marked and their attributes quantified on<br />

multiple occasions, it becomes possible to perform<br />

quantitative life history analysis (Brigham and Schwartz<br />

2003). One of the goals of life history analysis for plants<br />

of conservation concern is the formal inclusion of life<br />

history information into mathematical models that predict<br />

the probability of population persistence under different<br />

scenarios. The term used for the creation, execution,<br />

and interpretation of such models is population<br />

viability analysis (PVA; Beissinger and McCullough<br />

2002).<br />

Population viability analysis for plants is a relatively<br />

new area of research (Doak et al. 2002). One feature of<br />

plant life history that makes PVA difficult is the presence<br />

of a largely invisible life cycle stage, namely seeds<br />

in the soil seed bank. This paper deals both with assessing<br />

the need for including soil seed bank data in life history<br />

analysis for a particular plant species and tackling<br />

the problem of obtaining seed bank data when necessary.<br />

WHEN ARE SOIL SEED BANKS IMPORTANT?<br />

Doak and others (2002) provide an excellent discussion<br />

of problems associated with assumptions about<br />

seed banks in the context of PVA for plants. They first<br />

present the two extremes for plant life history strategy.<br />

At one extreme is a life history much like that of the<br />

vertebrates for which PVA was first developed. In these<br />

organisms, for example, giant sequoias, newborns are<br />

subject to high and variable mortality, the juvenile phase<br />

lasts a long time and is characterized by slow growth<br />

and increasing annual survival, and adults are longlived.<br />

<strong>Plant</strong>s with this life history usually produce seeds<br />

that have a very short tenure in the seed bank, less than<br />

a year. At the other extreme are plants that have short<br />

and risky juvenile stages and short life spans as adults,<br />

but whose seeds have high dormancy and high survivorship<br />

in the soil, resulting in a large, persistent seed bank.<br />

For plants in the former category, the seed bank can<br />

safely be ignored in PVA, because the transition from<br />

seed production to seedling takes place within a single<br />

year. But for plants that produce long-lived seeds, the<br />

dynamics of the seed bank can play a very important<br />

role in population biology, and false assumptions based<br />

on inadequate data can lead to major problems with the<br />

resulting PVA.<br />

Doak and others (2002) identify two plant life history<br />

attributes most critical for deciding how important accurate<br />

measurement of seed bank dynamics is likely to be.<br />

The first, as implied above, is plant longevity, and the<br />

second is the effect of a variable environment on adult<br />

survival and reproduction. A long and stable mature<br />

stage with multiple opportunities for reproduction is<br />

likely to minimize the importance of a persistent seed<br />

bank. On the other hand, short-lived plants which are at<br />

variable risk of mortality and/or reproductive failure as<br />

adults are likely to be associated with long-lived seeds<br />

and persistent seed banks. Different combinations of<br />

positions along these two gradients (life span and environmental<br />

risk to adults) result in different assessments<br />

of the importance of seed banks. Even relatively longlived<br />

shrubs may have persistent seed banks if there is a<br />

risk of catastrophic mortality to adult plants, for example<br />

through fire or epidemic disease. And even annuals<br />

may have relatively short-lived seeds if the chances of<br />

adult survival to seed production are high.<br />

Another source of clues about the importance of the<br />

persistent seed bank is in the germination behavior of<br />

46


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

the seeds in a laboratory setting. If the seeds are nondormant<br />

at dispersal and cannot readily be induced into<br />

dormancy, it is unlikely that they will be able to form a<br />

persistent seed bank under field conditions. Similarly, if<br />

the seeds are dormant at dispersal but lose dormancy in<br />

response to an environmental cue likely to be encountered<br />

before the optimum germination time within the<br />

year, they are also unlikely to form a persistent seed<br />

bank. This type of dormancy is called cue-responsive or<br />

predictive dormancy. It functions to time germination<br />

optimally within the year following production by allowing<br />

the seeds to sense their environment and respond<br />

appropriately. Many spring-germinating species in the<br />

temperate zone have this type of seed dormancy, with<br />

moist-chilling or cold stratification that simulates winter<br />

conditions as the cue. This prevents precocious germination<br />

the autumn following production but allows complete<br />

germination the following spring.<br />

Not all cue-responsive dormancy is associated with<br />

short-lived seed banks, however. Sometimes the cue is<br />

associated with an episodic event such as fire (Keeley<br />

1987), or tillage that exposes weed seeds to light<br />

(Baskin and Baskin 1985). Such seeds may persist in the<br />

soil for many years, but will germinate synchronously<br />

when the cue is received. These seeds germinate readily<br />

in response to a laboratory-administered cue. The trick<br />

is in recognizing that such a cue is unlikely to be encountered<br />

under field conditions within any particular<br />

year, so that seeds with this kind of cue-responsive dormancy<br />

form a persistent seed bank.<br />

The best laboratory clue that seeds of a species are<br />

likely to form a persistent seed bank under field conditions<br />

is the presence of cue-non-responsive dormancy.<br />

These seeds will germinate to only small percentages no<br />

matter what dormancy-breaking treatment is applied.<br />

Sometimes it is possible to determine what pretreatment<br />

is needed to make the seeds become responsive to a particular<br />

cue, but more often even this is very difficult.<br />

The individual seeds are programmed to come out of<br />

dormancy at different times over a protracted period,<br />

and there seems to be no way to speed or circumvent<br />

this process. Often the only way to break cue-nonresponsive<br />

dormancy is to resort to unnatural treatments<br />

like injuring the seed, and sometimes even these draconian<br />

measures fail to trigger germination.<br />

METHODOLOGIES FOR STUDYING SEED<br />

BANKS<br />

For many people, the most obvious way to begin a<br />

study of seed bank dynamics is to attempt to quantify<br />

the in situ seed bank. This involves taking seed bank<br />

samples from the field and somehow measuring the<br />

number of seeds these samples contain. A common<br />

measuring method involves spreading the soil samples<br />

out in shallow pans, applying water, and counting and<br />

removing germinants as they emerge. Usually the soil<br />

sample is turned multiple times to encourage subsequent<br />

flushes of germination and emergence, and the seed<br />

bank is considered depleted when no further emergence<br />

occurs. Obviously, this methodology involves numerous<br />

assumptions about the dormancy status of the seeds,<br />

because seeds that do not germinate and emerge as seedlings<br />

are not included in the quantification. Sometimes<br />

the seedling emergence methodology includes multiple<br />

cycles of application of dormancy-breaking cues, for<br />

example, moist-chilling, which increases the chances of<br />

complete germination. But for truly cue-non-responsive<br />

species, these methodologies are clearly inadequate.<br />

Another commonly applied method for quantifying<br />

the in situ seed bank involves flotation, usually using a<br />

chemical such as potassium carbonate at high molarity.<br />

This method has been shown to yield more seeds on<br />

average than the emergence (germination) methodology<br />

(Ishikawa-Goto and Tsuyuzaki 2004), but the extracted<br />

seeds are no longer viable. This inability to distinguish<br />

between viable and nonviable seeds and to evaluate seed<br />

dormancy status represents a major disadvantage to this<br />

method.<br />

A third method for quantifying in situ seed banks is<br />

rather labor-intensive, but avoids some of the problems<br />

associated with the previous two methods (Meyer et al.<br />

2007). The samples are dry-screened (or wet-screened,<br />

depending on the soil type) using sieve sizes that eliminate<br />

fine soil and large particles such as gravel and root<br />

chunks. The remaining fraction, which contains the<br />

seeds of interest, is hand-processed to remove the seeds,<br />

which can then be subjected to germination testing and/<br />

or viability evaluation. This method works best for medium<br />

to large seeds. Its accuracy is increased by inclusion<br />

of numerous small samples rather than fewer large<br />

samples, given an equal volume of sampled material.<br />

Sampling regime is a critical aspect of in situ seed<br />

bank evaluation, because the lateral distribution of seeds<br />

in soil is usually extremely heterogeneous. This means<br />

that stratified sampling regimes and often very large<br />

sample sizes are needed to get replicable data. It is<br />

highly advisable before launching into such a study to<br />

make sure that it is well-designed and will yield the desired<br />

information. In order to quantify the persistent<br />

seed bank, it is important to sample after germination is<br />

complete for the year but before any input of seed rain<br />

from current-year production, so that only seeds at least<br />

a year old will be sampled.<br />

Sampling the in situ seed bank cannot provide any<br />

information about seed bank persistence beyond a single<br />

year, because there is no way to know the age of seeds<br />

removed from the samples. Seeds from the previous<br />

production year could be the only ones present prior to<br />

dispersal of current-year seeds, or there could be an accumulation<br />

of seeds from an unknown number of prior<br />

47


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

production years. In situ sampling provides a quantification<br />

of the seed bank at a given point in time, but does<br />

not directly address seed bank dynamics.<br />

The method of choice for determining how long<br />

seeds can persist in the seed bank is the seed retrieval<br />

(seed burial or artificial seed bank) experiment. In this<br />

method, seeds of known age are introduced into the seed<br />

bank in a form that makes their subsequent retrieval and<br />

evaluation possible. This usually involves placing them<br />

in some sort of mesh bag that allows free passage of<br />

water and air, but not seeds. We usually use nylon mesh<br />

bags made from mosquito netting, which has a fine<br />

enough mesh size to contain even very small seeds. For<br />

larger seeds, fiberglass window screening works well.<br />

The nylon lasts for many years as long as it is covered<br />

with soil so that it cannot photo-degrade, and fiberglass<br />

screening can be placed directly on the surface.<br />

Some have criticized seed retrieval experiments because<br />

they place the seeds in an environment somewhat<br />

different from that experienced by seeds in the natural<br />

seed bank. An alternative method, or one that can be<br />

used in conjunction with a retrieval experiment, is an<br />

emergence experiment, where seeds of known age are<br />

planted in precise locations using a template or field<br />

marking system and monitored for emergence. For both<br />

retrieval and emergence experiments, it is important to<br />

include adequate replication and to use block designs to<br />

avoid localized effects that generate large experimental<br />

error.<br />

The retrieval method involves destructive sampling,<br />

so that it is necessary to include a different set of retrieval<br />

bags for each evaluation date. For rare plants,<br />

this can involve a prohibitively large number of seeds.<br />

But even yearly retrievals for three years, using a few<br />

hundred seeds, can provide valuable information that is<br />

difficult to obtain in any other way.<br />

Processing retrieval bags can be a zen experience,<br />

not suitable for those inclined to attention deficit. Particularly<br />

when the samples are wet after a germination<br />

event, it can take some time to untangle and quantify the<br />

germinants. Remaining ungerminated seeds are then<br />

incubated under controlled conditions and classified as<br />

germinable, dormant, or nonviable. With frequent sampling<br />

and adequate replication, it is possible to get a<br />

very good picture of the phenology of dormancy loss<br />

and germination, and also secondary dormancy induction,<br />

if it occurs.<br />

It is important to distinguish if possible between<br />

seeds that are lost from the seed bank through germination<br />

and those that are lost through pre-germination<br />

mortality. These two processes have potentially very<br />

different demographic consequences. Getting accurate<br />

information on field-germinated seeds requires retrieval<br />

soon after the germination event, and it can be difficult<br />

to anticipate the correct retrieval timing. Laboratory data<br />

can be helpful here. For example, if you know the seeds<br />

can germinate at near-freezing temperatures, it is reasonable<br />

to expect them to germinate during winter in<br />

places where the snow cover is deep enough to insulate<br />

the seed bed from freezing. This is a common strategy<br />

in environments where the soil dries quickly in the<br />

spring.<br />

PATTERNS OF SEED BANK DEPLETION<br />

Seed retrieval data can be examined graphically by<br />

plotting the percentage of initially viable seeds still present<br />

as viable seeds in the seed bank (actually, in the<br />

artificial seed bank) as a function of time in the field.<br />

For any given seed population, this will yield a characteristic<br />

seed depletion trajectory through time (fig. 1).<br />

For species that do not form persistent seed banks, this<br />

trajectory will essentially be a line that goes to a value<br />

of zero within a year of the initiation of the retrieval experiment<br />

(Figure 1a). Species with this type of seed depletion<br />

trajectory are said to have transient seed banks.<br />

In situ seed bank sampling after germination is complete<br />

but before current-year seed rain for species with transient<br />

seed banks is expected to yield no viable seeds.<br />

Basin wildrye (Leymus cinereus) is an example of a<br />

species with nondormant seeds and a transient seed<br />

bank (Meyer et al. 1995; Figure 2). Seed collections<br />

from three populations were placed in retrieval experiments<br />

in early fall at salt desert shrubland, sagebrush<br />

steppe, and mountain meadow sites. Germination did<br />

not take place in autumn, even though seeds were nondormant,<br />

because of their relatively long germination<br />

time. All the seeds germinated by the end of the following<br />

spring, regardless of population of origin or habitat<br />

at the seed retrieval site.<br />

Blackbrush (Coleogyne ramosissima) is another example<br />

of a species with a transient seed bank (Meyer<br />

and Pendleton 2004; Figure 3). Its seeds are dormant at<br />

dispersal in mid-summer but have cue-responsive dormancy,<br />

losing dormancy both in the dry state at high<br />

summer temperatures and during moist chilling. As a<br />

consequence, if winter rainfall is adequate, seeds germinate<br />

to high percentages during early spring. A few<br />

seeds may carry for a year over if the winter is unusually<br />

dry, but this is the exception rather than the rule.<br />

Seeds with cue-responsive dormancy that responds to<br />

a disturbance-associated cue show a similar pattern to<br />

those that form transient seed banks, except that there is<br />

an extended time delay prior to the seed bank-depleting<br />

germination event (Figure 1 b). These species are difficult<br />

to study in retrieval experiments, because the cue,<br />

such as fire or soil disturbance, is not only unpredictable<br />

but also would tend to destroy a retrieval experiment.<br />

Information on this type of seed bank depletion trajectory<br />

generally comes either from emergence experiments<br />

or from in situ seed bank studies that show abrupt<br />

48


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

100<br />

80<br />

A<br />

100<br />

80<br />

B<br />

60<br />

60<br />

Viable Seed Percentage<br />

40<br />

20<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

100<br />

80<br />

60<br />

40<br />

20<br />

C<br />

40<br />

20<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10<br />

100<br />

80<br />

60<br />

40<br />

20<br />

D<br />

0<br />

0<br />

0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10<br />

Retrieval Year<br />

Figure 1. Schematic seed bank depletion trajectories for: (A) a species with a transient seed bank and a depletion trajectory<br />

that reaches zero within one year, (B) a species with cue-responsive dormancy and a seed bank that persists<br />

until a specific dormancy-breaking cue is received, (C) a species with a short-persistent seed bank, a constant loss<br />

rate, and a negatively exponential depletion trajectory, and (D) a species with cue-non-responsive seeds and a linear<br />

seed bank depletion trajectory.<br />

decreases in seed density following receipt of the germination<br />

cue.<br />

In many herbaceous perennial species, most of the<br />

seeds are programmed to germinate during the first year<br />

following production, but some possess a mechanism<br />

permitting carryover for at least a year, even when conditions<br />

for dormancy release and germination the first<br />

year are optimal. These species are said to have shortpersistent<br />

seed banks. This germination response pattern<br />

results in a seed depletion trajectory that is essentially<br />

negatively exponential (Figure 1c). If rate of loss of a<br />

cohort of seeds in the seed bank is constant across years,<br />

this is the type of seed depletion trajectory that will be<br />

generated. For example, if 80% of the seeds are lost the<br />

first year, 80% of the remaining 20% are lost the second<br />

year, and 80% of the remaining 4% are lost the third<br />

year, this would generate a negatively exponential loss<br />

trajectory.<br />

Lewis flax (Linum lewisii) is an example of a species<br />

that may exhibit a negatively exponential loss trajectory<br />

(Meyer and Kitchen 1994; Figure 4). Seed dormancy<br />

loss and germination phenology in this species and its<br />

close relative L. perenne (Euopean blue flax) also vary<br />

as a function of both population of origin and habitat at<br />

the seed retrieval site. In a two-year retrieval experiment<br />

at three sites, the “Appar” release of European blue flax<br />

had nondormant seeds that formed only a transient seed<br />

bank, regardless of retrieval site habitat. In contrast, the<br />

montane Strawberry seed collection of Lewis flax was<br />

largely dormant at the initiation of the retrieval experiment<br />

and required chilling to become nondormant<br />

(Figure 4). It lost dormancy and germinated completely<br />

by the end of the first spring at its site of origin in the<br />

mountains, exhibiting the transient seed bank pattern. It<br />

carried over a substantial fraction through the end of the<br />

second year at the foothill and especially the salt desert<br />

site. Seeds placed outside of their environmental context<br />

can show very different germination patterns than those<br />

placed in the habitat of origin. These seeds did not receive<br />

sufficient chilling to break dormancy at the drier<br />

sites and tended to form a persistent seed bank under<br />

those conditions.<br />

The foothill Provo Overlook seed collection of Lewis<br />

flax was nondormant at the initiation of the retrieval<br />

experiment, but contained a fraction that could be induced<br />

into secondary dormancy early during chilling the<br />

first year (Figure 4). This resulted in a divergence of<br />

seed sub-populations, so that a sizeable fraction germi-<br />

49


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

VIABLE SEED PERCENTAGE<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

PINTO COLLECTION<br />

9/20 10/2011/<strong>2012</strong>/20 1/20 2/20 3/20 4/20 5/20 6/20<br />

INDIAN CANYON COLLECTION<br />

9/20 10/2011/<strong>2012</strong>/20 1/20 2/20 3/20 4/20 5/20 6/20<br />

STRAWBERRY COLLECTION<br />

9/20 10/2011/<strong>2012</strong>/20 1/20 2/20 3/20 4/20 5/20 6/20<br />

DATE<br />

WHITEROCKS DESERT SITE<br />

POINT OF THE MOUNTANI FOOTHILL SITE<br />

STRAWBERRY MONTANE SITE<br />

Figure 2. Seed bank depletion trajectories for three collections<br />

of basin wildrye (Leymus cinereus) placed in<br />

retrieval experiments at montane, foothill, and salt desert<br />

sites. All seeds germinated within a year of initiation<br />

of retrieval experiments. Adapted from Meyer and<br />

others (1995).<br />

nated the first year in all three habitats, but 20-30% remained<br />

in dormancy through the spring and carried over<br />

to the second year, generating the characteristic negatively<br />

exponential pattern. Three additional years of retrieval<br />

data for the Provo Overlook collection at these<br />

sites showed continued gradual decline in viable seed<br />

numbers each year (Meyer unpublished data).<br />

Seed bank depletion for shadscale (Atriplex confertifolia)<br />

exhibited a pattern somewhat similar to that of<br />

Lewis flax, but in this case the pattern resulted in a more<br />

persistent seed bank (Figure 5). There was no germination<br />

at all during the first year in the field, but a sizeable<br />

germination pulse occurred the second spring after retrieval<br />

initiation. During eight subsequent years in the<br />

field, the size of the remaining viable seed fraction<br />

slowly diminished until it was at or near zero for most<br />

collections (Figure 5; Meyer unpublished data). Laboratory<br />

germination experiments help to explain this pattern.<br />

Shadscale seeds are dormant at dispersal and tend<br />

to be nonresponsive to the chilling cue that triggers<br />

them to emerge at the correct time for establishment in<br />

early spring (Meyer et al. 1998). They must after-ripen<br />

in the dry state to become responsive to the chilling cue,<br />

and the rate of increase in the chilling-responsive fraction<br />

is an exponential function of temperature during<br />

dry storage (Garvin and Meyer 2003). This delays germination<br />

for at least a year, because the high summer<br />

temperatures needed for after-ripening are first experienced<br />

only after the first winter in this autumn-ripening<br />

species. Each summer another fraction becomes chilling-responsive,<br />

and this fraction is able to germinate the<br />

following spring if its chilling requirement is met. The<br />

resulting depletion trajectory is like a slow-motion version<br />

of the negative exponential pattern for species with<br />

short-lived seed banks. Once the first pulse of germination<br />

is complete, the trajectory becomes essentially linear<br />

through time.<br />

In contrast to the slow-exponential depletion trajectory<br />

seen for shadscale, species with truly cue nonresponsive<br />

dormancy tend to exhibit a linear depletion<br />

trajectory, with no large germination pulse in any one<br />

year and certainly not in the early years (Figure 1d).<br />

Hard-seeded legumes like <strong>Utah</strong> ladyfinger milkvetch<br />

(Astragalus utahensis) are typical of this group. Hardseededness<br />

refers to the inability of a seed to imbibe<br />

water because of physical barriers imposed by the seed<br />

coat or endocarp. It is often possible to trigger synchronous<br />

germination in hard-seeded species by injuring the<br />

seed coat, and it has been thought that such scarification<br />

must be a part of the natural dormancy-breaking regimen.<br />

Under field conditions, however, hard-seededness<br />

seems to be very gradually lost through time without<br />

any specific scarification mechanism. Because the seeds<br />

lose hard-seededness at different rates, this functions to<br />

spread germination over many years.<br />

50


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

PERCENTAGE<br />

OF INITIALLY VIABLE SEEDS<br />

100<br />

75<br />

50<br />

25<br />

0<br />

7-91 10-91 1-92 4-92 7-92<br />

RETRIEVAL DATE<br />

% DORMANT SEEDS<br />

% VIABLE UNGERMINATED SEEDS<br />

% VIABLE + GERMINATED SEEDS<br />

Figure 3. Change through time in the dormant seed fraction, the viable seed fraction, and the viable plus germinated<br />

seed fraction for blackbrush (Coleogyne ramosissima) seeds placed in a retrieval experiment in the habitat of origin at<br />

Arches National Park. All seeds germinated within a year of initiation of the experiment soon after dispersal.<br />

Adapted from Meyer and Pendleton (2004).<br />

Figure 4. Change through time during a two-year period for dormant seed percentage, viable ungerminated seed percentage,<br />

and viable plus germinated seed percentage for three collections of Linum placed in retrieval experiments at<br />

montane (Strawberry), foothill (Hobble Creek), and salt desert (Rush Valley) study sites. Adapted from Meyer and<br />

Kitchen (1994).<br />

51


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

PERCENTAGE (BASED ON INITIALLY VIABLE SEEDS)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

A<br />

8-COLLECTION MEAN<br />

11/91 3/92 7/92 11/92 3/93 7/93 11/93 3/94 7/94 11/94 3/95 7/95 11/95 3/96 7/96<br />

B<br />

PIPE<br />

SPRINGS<br />

11/91 3/92 7/92 11/92 3/93 7/93 11/93 3/94 7/94 11/94 3/95 7/95 11/95 3/96 7/96<br />

C<br />

NORTH OF MILFORD<br />

11/91 3/92 7/92 11/92 3/93 7/93 11/93 3/94 7/94 11/94 3/95 7/95 11/95 3/96 7/96<br />

RETRIEVAL DATE<br />

DORMANT NOT CHILL-RESPONSIVE<br />

DORMANT<br />

DORMANT+GERMINABLE<br />

DORMANT+GERMINABLE+FIELD-GERMINATED<br />

Figure 5. Patterns of change in the precentage of non-chill-responsive dormant seeds, chill-responsive dormant<br />

seeds, dormant plus germinable seeds and viable plus germinated seeds over a five-year period for: (A) The mean of<br />

eight seed collections, (B) the Pipe Springs seed collection, and (C) the north of Milford seed collection. Adapted<br />

from Meyer and others (1998).<br />

52


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Two or three years of retrieval data are usually sufficient<br />

to determine whether a species will show a negatively<br />

exponential or a linear depletion trajectory. When<br />

<strong>Utah</strong> ladyfinger milkvetch seeds were placed into retrieval<br />

experiments at three contrasting sites, depletion<br />

trajectories were clearly linear at all three sites (Figure<br />

6). The slope of the depletion trajectory was quite similar<br />

across habitats and showed no clear pattern as a<br />

function of habitat, indicating that rate of loss of hardseededness<br />

was not tightly tied to environmental conditions.<br />

For A. utahensis, regression equations based on the<br />

first two years of retrieval data at each site were able to<br />

predict the approximate size of the remaining fraction in<br />

the subsequent four years. These equations were also<br />

used to estimate maximum longevity of this seed population<br />

in the seed bank, which was about 14 years at the<br />

montane site, 10 years at the foothill site, and 12 years<br />

at the salt desert site. Including the later retrievals in<br />

these regressions did not change them significantly,<br />

even though the fit of the lines for these later dates was<br />

not as good. This retrieval experiment had only two replications<br />

per retrieval date, resulting in considerable error<br />

in the estimate of hard-seededness, especially in later<br />

years, when values dropped far below 100%. The regression<br />

equation may be the best indicator of actual<br />

rate of seed bank depletion under this scenario.<br />

100<br />

Rush Valley - Desert Study Site<br />

80<br />

60<br />

40<br />

Percentage of Initially Viable Seeds<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

Hard Seeds<br />

Hard + Germinable Seeds<br />

9-90 3-91 9-91 3-92 9-92 3-93 9-93 3-94 9-94 3-95 9-95 3-96 9-96<br />

Hobble Creek - Foothill Study Site<br />

9-90 3-91 9-91 3-92 9-92 3-93 9-93 3-94 9-94 3-95 9-95 3-96 9-96<br />

Strawberry - Montane Study Site<br />

60<br />

40<br />

20<br />

0<br />

9-90 3-91 9-91 3-92 9-92 3-93 9-93 3-94 9-94 3-95 9-95 3-96 9-96<br />

Date<br />

Figure 6. Patterns of change over a six-year period in the percentage of hard seeds and hard plus germinable seeds<br />

(total viable seeds) for a collection of <strong>Utah</strong> ladyfinger milkvetch (Astragalus utahensis) placed in seed retrieval experiments<br />

at three sites. Regression lines are fit to total viable seed percentage values at each site based on the first<br />

two years of retrieval (Rush Valley site: Percentage of viable seeds = -0.022 (days)+100.4, d.f.=13, R 2 = 0.794,<br />

P


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

All of the species discussed so far are common plants<br />

that were included in seed bank studies in order to understand<br />

their establishment ecology, rather than to provide<br />

seed bank data to inform PVA. There are very few<br />

long-term seed retrieval studies for rare plants. The<br />

Snake River Plains endemic Lepidium papilliferum is<br />

one of the few rare species whose seed bank dynamics<br />

have been included in PVA (Meyer et al. 2005, 2006).<br />

This spring ephemeral species has cue non-responsive<br />

seeds that show the characteristic linear decrease as a<br />

function of time in the proportion of initially viable<br />

seeds remaining in the seed bank (Figure 7; Meyer et al.<br />

2005). There was little or no germination during the first<br />

two years of this eleven-year retrieval study, so that at<br />

least three years of retrieval data would have been<br />

needed to estimate the slope of seed bank depletion. The<br />

maximum longevity in soil for two seed collections in-<br />

cluded in the study was estimated to be 12 years. Seeds<br />

of this species can be induced to germinate by piercing<br />

imbibed seeds and subjecting them to 2-4 weeks of<br />

moist chilling. Development of this technique has facilitated<br />

greenhouse production of seeds for reintroduction<br />

experiments, but seems to shed little light on how the<br />

seeds gradually become nondormant in the field.<br />

Rare plants can be expected to have seeds that run<br />

the gamut of germination response patterns and associated<br />

seed bank depletion trajectories. As discussed<br />

above, it is frequently important to investigate this aspect<br />

of rare plant population biology, especially when<br />

the goal is to understand the likely future status for a<br />

population or species. The information obtained from<br />

seed bank studies can also be critical for planning management<br />

and mitigation activities.<br />

% OF INITIALLY VIABLE SEEDS<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1991 SEEDS<br />

8/92 8/93 8/94 8/95 8/96 8/97 8/98 8/99 8/2000 8/2001 8/2002 8/2003<br />

1992 SEEDS<br />

DORMANT<br />

DORMANT PLUS GERMINABLE<br />

DORMANT PLUS GERMINABLE PLUS GERMINATED<br />

TOTAL I NITIALLY VIABLE<br />

8/92 8/93 8/94 8/95 8/96 8/97 8/98 8/99 8/2000 8/2001 8/2002 8/2003<br />

DATE<br />

Figure 7. Patterns of change through time in dormant seed percentage, dormant plus germinable seed percentage, and<br />

viable plus germinated seed percentage over an eleven-year period in a seed retrieval experiment with two collections<br />

of Lepidium papilliferum. The difference between viable plus germinated seed percentage and initially viable seed<br />

percentage represents seeds that lost viability prior to germinating. Adapted from Meyer and others (2005).<br />

54


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

LITERATURE CITED<br />

Baskin, J.M., and C.C. Baskin. 1985. The annual<br />

cycle in buried weed seeds: a continuum. BioScience<br />

35:492-998.<br />

Beissinger, S.R and D.R. McCullough (eds). 2002.<br />

Population Viability Analysis. University of Chicago<br />

Press, Chicago, Illinois. 577 p.<br />

Brigham, C.A, and M.W. Schwartz, eds. 2003. Population<br />

viability in plants. Conservation, management<br />

and modeling of rare plants. Ecological Studies,<br />

Vol.165. Springer, Berlin. 362 p.<br />

Doak, D.F., D. Thompson, and E.S. Jules. 2002.<br />

Population viability analysis for plants: understanding<br />

the demographic consequences of seed banks for population<br />

health. p.312-337. In: Beissinger, S. R and D. R.<br />

McCullough (eds). Population Viability Analysis. University<br />

of Chicago Press, Chicago, Illinois.<br />

Garvin, S.C., and S.E. Meyer. 2003. Multiple mechanisms<br />

of seed dormancy regulation in shadscale<br />

(Atriplex confertifolia: Chenopodiaceae). Canadian<br />

Journal of Botany 81:301-610.<br />

Ishikawa-Goto, M., and S. Tsuyuzaki. 2004. Methods<br />

of estimating seed banks with reference to longterm<br />

seed burial. Journal of <strong>Plant</strong> Research 117: 245-<br />

248.<br />

Keeley, J.E. 1987. Role of fire in seed germination of<br />

woody taxa in California chaparral. Ecology 68: 434-<br />

443.<br />

Meyer, S.E., J. Beckstead, P.S. Allen, and H. Pullman.<br />

1995. Germination ecophysiology of Leymus<br />

cinereus (Poaceae). International Journal of <strong>Plant</strong> Sciences<br />

156: 206-215.<br />

Meyer, S.E., S.L. Carlson, and S.C. Garvin. 1998.<br />

Seed germination regulation and field seed bank carryover<br />

in shadscale (Atriplex confertifolia: Chenopodiaceae).<br />

Journal of Arid Environments 38: 255-267.<br />

Meyer, S.E. and S.G. Kitchen. 1994. Life history<br />

variation in blue flax (Linum perenne: Linaceae): Seed<br />

germination phenology. American Journal of Botany 81:<br />

528-535.<br />

Meyer, S.E. and B.K. Pendleton. 2005. Factors affecting<br />

seed germination and seedling establishment of a<br />

long-lived desert shrub (Coleogyne ramosissima:<br />

Rosaceae). <strong>Plant</strong> Ecology 178:171–187.<br />

Meyer, S.E., D. Quinney, D.L. Nelson, and J.<br />

Weaver. 2007. Impact of the pathogen Pyrenophora<br />

semeniperda on Bromus tectorum seedbank dynamics in<br />

North American cold deserts. Weed Research 47: 54–<br />

62.<br />

Meyer , S.E, D. Quinney, and J. Weaver. 2005. A life<br />

history study of the Snake River Plains endemic<br />

Lepidium papilliferum (Brassicaceae). Western North<br />

American Naturalist 65:11-23.<br />

Meyer, S.E., D. Quinney, and J. Weaver. 2006. A<br />

stochastic population model for Lepidium papilliferum<br />

(Brassicaceae), a rare desert ephemeral with a persistent<br />

seed bank. American Journal of Botany. 93: 891–902.<br />

55


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

East Meets West: Rare Desert Alliums in Arizona<br />

John L. Anderson,<br />

US Bureau of Land Management, Phoenix, AZ, retired<br />

Abstract. Two previously poorly known desert species of Allium in Arizona, A. bigelovii and A. parishii, were investigated<br />

to determine their actual rarity and conservation status. Each approached Arizona from different deserts<br />

and opposite directions: Allium bigelovii from the Chihuahuan Desert in New Mexico and Allium parishii from the<br />

Mohave Desert in California. The status of Allium bigelovii was made difficult to understand by the many misidentifications<br />

of herbarium specimens. A review of herbarium specimens and field visits determined that Allium bigelovii is<br />

a Chihuahuan Desert species that enters southeastern Arizona and is disjunct farther west into the Sonoran Desert on<br />

unusual lacustrine soils. Allium parishii was only known from historic collections in two mountain ranges in western<br />

Arizona. These historic locations were relocated. Although these two Allium species normally occur nearly 500 km<br />

apart and in different deserts, they approach within less than 100 km in the Sonoran Desert of western Arizona.<br />

There are 96 species of Allium L. in North America<br />

north of Mexico with thirteen species in Arizona<br />

(McNeal and Jacobsen 2002). The taxonomy of Allium<br />

in Arizona is largely unchanged since the monograph of<br />

Ownbey (1947). Three of the species recognized there<br />

are now treated as varieties: Allium nevadense S. Wats.<br />

var. cristatum (S. Wats.) Ownbey as A. atrorubens S.<br />

Wats. var. cristatum (S. Wats.) D. McNeal; A. palmeri<br />

Wats. as A. bisceptrum S. Wats. var. palmeri (Wats.)<br />

Cronquist; and A. rubrum Osterhout as A. geyeri S.<br />

Wats. var. tenerum Jones. The thirteen Arizona species<br />

are rather equally divided through the diverse Arizona<br />

habitats of deserts and mountains, aridlands and wetlands,<br />

low and high elevations, and northern and southern<br />

floristic affinities. Two of the desert species, Allium<br />

bigelovii S. Wats. (Figure 1) and A. parishii S. Wats.<br />

(Figure 2), have been poorly known in Arizona, but for<br />

different reasons. The taxonomic identity of the former,<br />

and consequently its real range and habitat in Arizona,<br />

has been confused by the many misidentifications of<br />

herbarium specimens; and, the geographic status of the<br />

latter in Arizona was made unclear by its few historic<br />

records in the state.<br />

The range of Allium bigelovii was described by Ownbey<br />

(1947) as “…southwestern New Mexico, northwestward<br />

across central Arizona to Mohave County.” In Arizona<br />

he cited five historic collections: (isotype, Palmer<br />

532 NY 1876 (Figure 3); Rusby 839 NY 1883; Crooks<br />

et al ARIZ 1939; Crooks & Darrow ARIZ 1938; and<br />

Benson & Darrow POM 1941), all from central Arizona<br />

(Figure 4). The type collection of A. bigelovii is from<br />

Cook’s Springs (Bigelow s.n.) in southwestern New<br />

Mexico (Watson 1871). Several collections at RSA<br />

(Eastwood 8276, Greene s.n.., Jones s.n.) and at NY<br />

(Greene s.n., Rusby s.n.., and Holmgren 6891) document<br />

its historic occurrence in southwestern New Mexico.<br />

Sivinski (2003) described the habitat and range of<br />

Figure 1. Allium bigelovii.<br />

Figure 2. Allium parishii.<br />

56


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 4. Map of the Arizona distribution of Allium<br />

bigelovii based on Ownbey (1947).<br />

Figure 3. First collection of Allium bigelovii in Arizona,<br />

Palmer 532 Walnut Grove, Yavapai County.<br />

Allium bigelovii in New Mexico as “…a desert species<br />

of southwestern New Mexico”, therefore, a Chihuahuan<br />

Desert species. At the start of this study twenty six collections<br />

at Arizona herbaria (ARIZ, ASU, ATC, Grand<br />

Canyon National Park (GC), Museum of Northern Arizona<br />

(MNA), and BLM Safford [BLMS]), had been<br />

identified or annotated as A. bigelovii (Table 1). Based<br />

upon these annotations, A. bigelovii was widely variable<br />

in habitat from desert to chaparral, grassland, oak woodland,<br />

pinyon-juniper woodland, and ponderosa pine and<br />

wide ranging geographically from southeastern Arizona<br />

westward beyond Wickenburg into the Sonoran Desert<br />

and northwestward to the Hopi lands and the Grand<br />

Canyon on the Colorado Plateau (Figure 5). This situation<br />

made its identity as a species and its natural habitat<br />

in Arizona puzzling.<br />

By examining these collections, I determined that<br />

over half had been misidentified and only twelve were<br />

actually Allium bigelovii (Table 1). The misidentified<br />

collections were annotated by me to several other species,<br />

the majority (8) as A. bisceptrum var. palmeri,<br />

three as A. macropetalum, two as A. acuminatum, and<br />

one as A. atrorubens var. cristatum (Table 1). The large<br />

number of misidentifications of Allium bigelovii as A.<br />

bisceptrum var. palmeri was due to their pairing in the<br />

key in Kearney and Peebles (1960) and the vague species<br />

differentiation based on qualitative characters there.<br />

To make species determinations I used the two species’<br />

descriptions in Ownbey (1947) that included quantitative<br />

morphological characters as well as information on<br />

herbarium labels of habitat and geographic data. The<br />

basis for species definition was thus an evolutionary<br />

combination of morphology and ecology, a species’<br />

physical characters, and the niche a taxonomic entity<br />

occupies in nature. The refinement of the identity of A.<br />

bigelovii in Arizona combined with a knowledge of its<br />

habitat and range in New Mexico demonstrated that A.<br />

bigelovii is a Chihuahuan Desert species (Figure 6) that<br />

extends from southwestern New Mexico into southeastern<br />

Arizona (Figure 7) (Gunder AZ930-8 ASU; Lunt 6<br />

BLMS).<br />

The localities of the remaining Arizona collections of<br />

Allium bigelovii followed an interesting disjunct pattern<br />

of distribution that extended the range of A. bigelovii<br />

discontinuously on Mid-Late Tertiary lacustrine deposits<br />

across central Arizona. This relictual “steppingstone”<br />

pattern had been documented by me (Anderson<br />

1996) for many species of various floristic affinities including<br />

other Chihuahuan Desert species: Anulocaulis<br />

leisolenus (Torrey) Standl., Polygala scoparioides Chodat.,<br />

and Thamnosma texana (A. Gray) Torrey. The<br />

lacustrine deposits (Nations et al. 1982) containing A.<br />

57


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. Author annotations of specimens at Arizona herbaria: ARIZ, AU, ATC, BLM Safford (BLMS),<br />

Grand Canyon National Park (GC), and Museum of Northern Arizona (MNA), previously identified as Allium<br />

bigelovii .<br />

Allium bigelovii Allium acuminatum Allium atrorubens<br />

var cristatum<br />

ARIZ: Crooks s.n.<br />

1938; Crooks s.n.<br />

1939; Darrow 10906.<br />

ASU: Butterwick<br />

4504, 6165; Gunder<br />

s.n. ATC: Kierstad 80<br />

-1; Morefield 1324;<br />

Llambreschte 30.<br />

BLMS: Lunt s.n.<br />

MNA: Haskell &<br />

Deaver 2449; Wetherill<br />

s.n.<br />

ARIZ: Reichenbacher<br />

1391. ASU: Lehto<br />

21378.<br />

Allium bisceptrum<br />

var. palmeri<br />

GC: Stochert s.n. ARIZ: Fishbein 304;<br />

Warren 248. ATC:<br />

Kasch s.n. 1979, s.n.<br />

1980; Llamphear s.n.;<br />

Reese s.n. BLMS:<br />

Bingham 3225. MNA:<br />

Kewanwytewa s.n.<br />

Allium macropetalum<br />

ARIZ: Wright s.n.<br />

ASU: Duran s.n.;<br />

Toleman 6-N.<br />

Figure 5. Map of the Arizona distribution of Allium<br />

bigelovii based on twenty six collections identified as<br />

Allium bigelovii at Arizona herbaria.<br />

Figure 6. Revised map of Arizona distribution of Allium<br />

bigelovii based on author’s annotations of collections at<br />

Arizona herbaria.<br />

58


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 7. Chihuahuan Desert habitat of Allium bigelovii<br />

in Greenlee County, Arizona.<br />

bigelovii include, from southeast to northwest, Tonto<br />

Basin (Crooks s.n. 1939 ARIZ); Verde Formation -<br />

Verde Valley (Morefield 1324 ATC; Lambreschte 30<br />

ATC; Haskell & Deaver 2449 MNA; Wetherill s.n.<br />

MNA); Rock Springs beds - Table Mountain (Kierstead<br />

80-1 ATC); Milk Creek beds - Walnut Grove (Palmer<br />

532 NY); Burro Creek (Crooks s.n. 1938 ARIZ; Darrow<br />

10906 ARIZ; Butterwick 4504 ASU; Anderson 2008-07<br />

ASU; and Chapin Wash Formation - Anderson Mine<br />

(Otton 1981) (Butterwick 6165 ASU; Anderson 2008-03<br />

ASU [Appendix 1]). These disjunct localities brought<br />

the Chihuahuan Desert species, Allium bigelovii, into<br />

the Sonoran Desert of west central Arizona as far west<br />

as Burro Creek, Mohave County (Figures 8, 9). Allium<br />

bigelovii is rarer in Arizona than previously thought. It<br />

is now known from approximately eight to ten localities<br />

(some collections from the Verde Valley have vague<br />

locality data on the herbarium labels). Because it occurs<br />

throughout southwestern New Mexico and is “…occasionally<br />

abundant…” there (Sivinski 2003), A. bigelovii<br />

is not a rare species overall.<br />

Allium parishii is a rare Mohave Desert species from<br />

California with peripheral localities in western Arizona<br />

at the eastern edge of its range (Figure 10). The type<br />

collection of A. parishii is from Cushenbury Springs,<br />

San Bernardino County, CA (S. B. Parish 1344 NY)<br />

(Watson 1882). A recent review of Allium parishii by<br />

White (2005) documented its current range and status.<br />

In California it primarily occurs in the San Bernardino<br />

Mountains (San Bernardino County), Little San Bernardino<br />

Mountains (Riverside County) and eastward into<br />

Joshua Tree National Park (JTNP). White (2005) recommends<br />

CNPS List 1B status. Its range in Joshua Tree<br />

National Park has recently been expanded by T. La<br />

Doux, botanist at JTNP (pers. comm. 2008).<br />

In Arizona, Allium parishii was once known only<br />

from an historic collection by Marcus Jones in 1903<br />

Figure 8. Sonoran Desert habitat of Allium bigelovii on<br />

lacustrine habitat at Burro Creek, Mohave County, Arizona,<br />

(westernmost occurrence).<br />

Figure 9. Close up of Allium bigelovii at Burro Creek.<br />

(Jones s.n., POM) from the Chemehuevi Mountains<br />

(Figure 11) which are now identified as the Mohave<br />

Mountains just east of Lake Havasu City, Mohave<br />

County. In 2005, students from Northern Arizona University<br />

made Allium collections in the Mohave Mountains.<br />

These collections were originally labeled as Allium<br />

atrorubens (Aamodt 9 ATC) and A. nevadense<br />

(Dow 13 ATC), but I subsequently identified them as A.<br />

parishii. These collections of Allium parishii were thus<br />

from the same mountain range as the historic Jones collection,<br />

but they included specific GPS locality data. In<br />

2008 I relocated A. parishii in this area (Figure 12) and<br />

recorded habitat data, associated species, and GPS locations<br />

(Anderson 2008-05 ASU [Appendix 1]). Interestingly,<br />

the two sites I recorded were only a mile apart but<br />

the plants grew on soils from different geological substrates:<br />

granite at Scott’s Well (Figure 13) and metamorphic<br />

gneiss near Arrastra Well (Figure 14). Also, the<br />

former site contained a diverse Sonoran/Mohave desert<br />

59


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 12. Allium parishii in the Mohave Mountains.<br />

Figure 10. Map of Allium parishii distribution in Arizona.<br />

Figure 13. Mohave Mountains habitat, Mohave County,<br />

Arizona, of Allium parishii with diverse Sonoran/<br />

Mohave Desert shrubs on granitic soils.<br />

Figure 11. Jones s.n. 1903 collection of Allium parishii<br />

from the Chemehuevi Mountains.<br />

Figure 14. Mohave Mountains habitat of Allium parishii<br />

in low diversity burned habitat on metamorphic gneiss<br />

soils.<br />

60


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

shrub community whereas the latter site had much less<br />

shrub diversity and a dense growth of Bromus rubens L.<br />

(red brome) (Figure 15) due to a wildfire as evidenced<br />

by old burn scars on Yucca brevifolia stumps.<br />

More recent collections have been made from the<br />

Kofa Mounains near Quartzite, in La Paz/Yuma Counties,<br />

50 miles south of the Mohave Mountains, beginning<br />

in 1937 (Nichols s.n. ARIZ) and continuing in<br />

1960 (Monson s.n. ARIZ) and 1976 (Irwin 33 ARIZ).<br />

Also in 2005, Allium parishii was relocated in the Kofa<br />

Mountains by Karen Reichhardt from the BLM Yuma<br />

Field Office (Reichhardt 2005-100 ARIZ). The Kofa<br />

Mountains are a well known locality for many relict<br />

species including Washingtonia filifera Wendl. and Berberis<br />

harrisoniana Kearney and Peebles. I visited this<br />

site and recorded the habitat, associated species, and<br />

GPS location (Anderson 2008-04 ASU [Appendix 1]).<br />

In the Kofa Mountains A. parishii occurs on yet another<br />

edaphic habitat: volcanic soils derived from andesite, in<br />

a diverse Sonoran Desert shrub community (Figure 16).<br />

The Kofa Mountains site showed no evidence of wildfire<br />

and contained little Bromus rubens (Figure 17). A.<br />

parishii seems to be a resilient species due to the ecological<br />

variability in its geological and edaphic habitats<br />

and its ability to survive the increasing wildfire frequency<br />

scourging the Sonoran and Mohave Deserts due<br />

to the growing presence of Bromus rubens.<br />

The biogeographical effect of past climatic and geological<br />

history on plant species migrations can result in<br />

unusual patterns of distribution. The research presented<br />

here has demonstrated the surprising geographic proximity<br />

of these two desert species of Allium previously<br />

known primarily from opposite sides of Arizona (Figure<br />

18). The accurate delineation of Allium bigelovii as a<br />

Chihuahuan Desert species with a disjunct population as<br />

far west as Burro Creek and the rediscovery of A.<br />

Figure 16. Kofa Mountains habitat, La Paz County, Arizona,<br />

of Allium parishii on volcanic andesitic soil.<br />

parishii, a California Mohave Desert species, at the<br />

eastern edge of its range in the Mohave Mountains,<br />

brings these two species, usually 500 km apart, to within<br />

less than 100 km of each other in the Sonoran Desert of<br />

Mohave County, Arizona.<br />

ACKNOWLEDGMENTS<br />

I want to thank Edward Gilbert and Elizabeth Makings<br />

from Arizona State University for help in making<br />

maps and J. Mark Porter from Claremont College, Rancho<br />

Santa Ana Botanic Garden, for help in making<br />

Figure 15. Mohave Mountains burned habitat of Allium<br />

parishii with thick cover of Bromus rubens.<br />

Figure 17. Kofa Mountains habitat of Allium parishii<br />

showing lack of Bromus rubens.<br />

61


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Sivinski, R.C. 2003. Annotated checklist of the genus<br />

Allium (Liliaceae) in New Mexico. The New Mexico<br />

Botanist 27: 1-6.<br />

Watson, S. 1871, U. S. Geol, Expl. 40 th Par. 487, pl<br />

38, fig 8,9.<br />

Watson, S. 1882. Proc. Amer. Acad. Arts 17: 380.<br />

White, S.D. 2005. Status review: Allium parishii<br />

Watson. Crossosoma 30: 5-16.<br />

APPENDIX 1<br />

All voucher specimens are deposited at Arizona State<br />

University (ASU).<br />

Figure 18. Map of Allium bigelovii and A. parishii<br />

showing distributional proximity in Mohave County,<br />

Arizona.<br />

maps and figures. Christine Bates, Karen Reichhardt,<br />

and Dick Zeller provided field assistance. Dr. D. Mc-<br />

Neal, Dr. D. Pinkava, and Mr. R. Sivinski reviewed the<br />

manuscript.<br />

LITERATURE CITED<br />

Anderson, J.L. 1996. Floristic patterns on late Tertiary<br />

lacustrine deposits in the Arizona Sonoran Desert.<br />

Madrono 43: 255-272.<br />

Kearney, T.H. and R.H. Peebles. 1960. Arizona<br />

Flora, second edition, with supplement, by J.T. Howell<br />

and E. McClintock. Univ. of Calif. Press, Berkeley, CA.<br />

McNeal, D.W. and T.D. Jacobsen. 2002. Allium. In:<br />

Flora of North America Editorial Committee, eds.<br />

1993+. Flora of North America North of Mexico. 12+<br />

vols. New York and Oxford. Vol. 26: 224-261.<br />

Nations, J.D., R.H. Hevley, D.W. Blinn, and J.J.<br />

Landye. 1982. Location and chronology of Tertiary<br />

sedimentary deposits in Arizona: a review. Pp. 107-122<br />

in RV. Ingersoll and M. O. Woodburne (eds.), Cenozoic<br />

nonmarine deposits of California and Arizona. Pacific<br />

Section, <strong>Society</strong> of Economic Paleontologists and<br />

Mineralogists.<br />

Otton, J.K. 1981. Geology and enesis of the Anderson<br />

mine, a carbonaceous lacustrine uranium deposit,<br />

western Arizona: a summary report. U. S. Geological<br />

Survey Open- File Report 81-780.<br />

Ownbey, M. 1947. The genus Allium in Arizona.<br />

Research Studies of the State College of Washington<br />

15:211-232.<br />

Allium bigelovii S. Wats.<br />

Arizona: Yavapai Co.: Anderson Mine site, late Tertiary<br />

lacustrine outcrop; with Canotia holocantha, Ambrosia<br />

dumosus, Fouquieria splendens, Nolina bigelovii, Yucca<br />

brevifolia, Pleuraphis rigida, Calochortus flexuosus;<br />

Tiquilia canescens; Locally common; 12S 0290870<br />

3798207 1945 ft. John L. Anderson 2008-03, Apr 7,<br />

2008.<br />

Arizona: Mohave Co.: Burro Creek Cliffrose site, ca 2<br />

miles above Six Mile Crossing of Burro Creek; late Tertiary<br />

lacustrine outcrop; with Canotia holocantha, Polygala<br />

acanthoclada, Chrysothamnus nauseosus var.<br />

junceus, Ziziphus obtusifolia, Calochortus flexuosus,<br />

Pleuraphis rigida, Aristida purpurea, Dichelostemma<br />

pulchra; Locally common; 12S 0283300 3828471<br />

2436 ft. John L. Anderson 2008-07, Apr 24, 2008.<br />

Allium parishii S. Wats.<br />

Arizona: La Paz Co.: Kofa Mts., High Tank Seven side<br />

canyon off of Burro Canyon; north-facing andesite hillside;<br />

with Simmondsia chinensis, Bernardia incana,<br />

Eriogonum fasciculatum, Fouquieria splendens,<br />

Ephedra aspera, Canotia holocantha, Viguieria deltoides,<br />

Acacia greggii, Krameria grayi, Gallium stellatum,<br />

Xylorhiza tortifolia, Pleuraphis rigida, Stipa speciosa,<br />

Calochortus kennedyi, Dichelostemma pulchra,<br />

Opuntia chlorotica, Agave desertii; uncommon (ca 50<br />

plants on one acre surveyed); 11S 0777891 3698772<br />

2788 ft. John L. Anderson 2008-04, Apr 17, 2008.<br />

Arizona: Mohave Co.: Mohave Mts., side canyon with<br />

Scotts Well, ca ½ miles NE of Scotts Well; north-facing<br />

granite hillside; with Canotia holocantha, Nolina bigelovii,<br />

Eriogonum fasciculatum, Encelia virginensis, Galium<br />

stellatum, Acacia greggii, Viguieria deltoides, Lotus<br />

rigida, Ephedra aspera, Opuntia acanthacarpa, Thamnosma<br />

montana, Janusia gracilis, Xylorhiza tortifolia,<br />

Pleuraphis rigida, Stipa speciosa; uncommon (ca 100<br />

plants on ten acres surveyed); 11S 758164 3829072<br />

3220 ft; T14N R 18W S7 NENW; John L. Anderson<br />

2008-05, Apr 23, 2008.<br />

62


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Spatial Patterns of Endemic <strong>Plant</strong> Species of the Colorado Plateau<br />

Crystal M. Krause,<br />

Northern Arizona University, Biological Sciences, Flagstaff, AZ<br />

Abstract. The Colorado Plateau region supports one of the highest levels of endemism in the United States. Of the<br />

6,800 vascular plants of the region more than 300 are endemic. Endemic species may have a higher risk of extinction<br />

due to their restricted geographic range. This risk may be increased with climate change. To better understand the risk<br />

to endemics, ecological niche modeling can provide a better understanding of the dynamics of environmental factors<br />

on a species range. For the endemics of the Colorado Plateau, a changing climate may modify species range. But underlying<br />

factors such as substrate and specialized habitat will also play a role in how a species range may change. The<br />

focus of this study is to understand spatial patterns and factors that predict endemism and then model species potential<br />

distribution.<br />

The Colorado Plateau ecoregion supports one of the<br />

highest levels of endemism in North America, ranking<br />

in the top three ecoregions on the continent for the total<br />

number of endemics in all taxonomic groups (Ricketts<br />

et al 1999). The Colorado Plateau also has the highest<br />

rate of endemism in terms of actual numbers of species<br />

(Kartesz and Farsted 1999). The harsh, dry environment<br />

of the Colorado Plateau has historically placed intense<br />

environmental stress on the flora. Factors such as soil,<br />

climate, and water scarcity among others seem to limit<br />

the geographic range of many species. More than 300<br />

species of vascular plants on the Colorado Plateau are<br />

endemic. Many of the endemic plant species are edaphic<br />

endemics restricted to one soil type, but larger geographic<br />

patterns of plant endemism can also be seen in<br />

the Colorado Plateau’s “sky island” habitats. Other important<br />

areas are those below 2000m; these lower elevations<br />

have a greater number of endemics than higher<br />

elevations (Welsh 1978).<br />

The Colorado Plateau contains 122,805,655 acres of<br />

land, of these 3,622,942 acres (3 percent) are protected<br />

lands in National Parks and Monuments, another<br />

64,748,735 acres (52 percent) are federally owned<br />

(Figure 1). Land ownership is also unique for the Plateau<br />

with the third most federally controlled land per<br />

area of all other ecoregions. Protected areas of the Colorado<br />

Plateau have a pivotal role to play in enabling species<br />

and ecosystems to persist. Protected areas can remove<br />

or control many of the threatening processes such<br />

as habitat loss and fragmentation.<br />

The distribution of endemic plants in protected areas<br />

is not fully known and very little work has been completed<br />

in modeling distribution shifts in response to climate<br />

change. To better understand the complexity and<br />

variability climate change may have on the distribution<br />

of plants and animals, ecologists have recently developed<br />

the concept of computationally based Ecological<br />

Niche Models (ENMs) (Peterson, Soberon and Samcjez-<br />

Cordero 1999; Peterson and Vieglais 2001; Stockwell<br />

and Peters 1999). ENMs integrate a wide range of environmental<br />

data (including point location data) to define<br />

potential species habitats. The output of ENMs is a set<br />

of grids of potential habitat based on the co-occurrence<br />

of known species locations and various environmental<br />

conditions. Each grid is assessed for accuracy by comparing<br />

a set of reserved species locations to the predicted<br />

habitat distributions.<br />

The predicted ENM habitats can be projected onto<br />

past, current, and modeled future landscapes thus providing<br />

testable habitat conditions that can be compared<br />

to known conditions to assess model accuracy and improve<br />

environmental predictions (Peterson et al. 2002).<br />

Projection onto the current landscape indicates the present<br />

day geographic distribution of suitable conditions<br />

for these plant species - the species potential habitat.<br />

Comparing these projections to historical and current<br />

location data provides information on the changes that<br />

plants have made in their dispersal in response to disturbances<br />

and environmental changes in the recent past.<br />

Projecting the model onto future landscapes provides<br />

information of how climate change may affect the species<br />

distribution and dispersal ability.<br />

METHODS<br />

MaxEnt<br />

Data collection is the first step in modeling a species<br />

distribution; species location point records and environmental<br />

data are needed for the model. Location point<br />

records used in this study are from herbaria and Natural<br />

Heritage Programs from the Four Corners region, National<br />

Park Service and the Bureau of Land Management.<br />

Species with occurrence points from less than 10<br />

populations were excluded from modeling because prior<br />

studies have demonstrated that fewer than 10 populations<br />

is not meaningful without extensive habitat re-<br />

63


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 1: Colorado Plateau Study Area Map<br />

quirement data and climate envelope constraints that<br />

may not be available (Stockwell and Peterson 2002).<br />

Environmental data listed in Table 1 are from www.<br />

worldclim.org and the USGS Hydro1K digital elevation<br />

model data. These variables include precipitation, temperature,<br />

aspect and slope. Geologic layers are from the<br />

USGS 1995 soils data. The geologic layers were converted<br />

from polygon to raster and downscaled to 1 km<br />

to match the climate and elevation layers.<br />

The modeling technique used in this study was Maximum<br />

Entropy or MaxEnt. MaxEnt is a general-purpose<br />

machine learning method of ecological niche modeling<br />

(Phillips, Anderson and Schapire 2006). MaxEnt estimates<br />

a species probability distribution by finding the<br />

probability distribution of maximum entropy, subject to<br />

a set of constraints that represent the information about<br />

the species distribution (Phillips, Anderson and Schapire<br />

2006).<br />

An important factor for choosing MaxEnt was that it<br />

allows for the use of presence-only data and categorical<br />

variables. In addition, MaxEnt has been shown to perform<br />

better than other algorithms for modeling distributions<br />

with limited data points (Elith et al 2006 and Pear-<br />

son et al 2007). MaxEnt provides mechanisms to assess<br />

the relative importance of each independent variable,<br />

which provides a better understanding of range shifts<br />

due to climate change.<br />

To calibrate the models all location points were randomly<br />

divided into training (70 percent) and testing (30<br />

percent) datasets. To evaluate the accuracy of the model<br />

and each variable’s predictive power, the Receiver Operating<br />

Characteristic (ROC) curve (Hanley and McNeil<br />

1982) was used for both training and test data. The ROC<br />

curve represents the relationship between the percentage<br />

of presences correctly predicted (sensitivity) and one<br />

minus the percentage of the absences correctly predicted<br />

(specificity). The Area Under the Curve (AUC) measures<br />

the ability of the model to classify correctly a species<br />

as present or absent. AUC values can be interpreted<br />

as the probability that when a site with the species present<br />

and a site with the species absent are drawn at random,<br />

the former will have a higher predicted value than<br />

the latter. Following Araujo and Guisan (2006), a rough<br />

guide for classifying the model accuracy is: 0.5-<br />

0.6=insufficient, 0.6-0.7=poor, 0.7-0.8=average, 0.8-<br />

0.9=good and 0.9-1=excellent.<br />

64


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. Environmental Variables<br />

BIO 1-Annual Mean Temperature<br />

BIO 2-Mean Diurnal Range<br />

BIO 3-Isothermality<br />

BIO 4-Temperature Seasonality<br />

BIO 5-Maximum Temperature of Warmest Month<br />

BIO 6-Minimum Temperature of Coldest Month<br />

BIO 7-Temperature Annual Range<br />

BIO 8-Mean Temperature of Wettest Quarter<br />

BIO 9-Mean Temperature of Driest Quarter<br />

BIO 10-Mean Temperature of Warmest Quarter<br />

BIO 11-Mean Temperature of Coldest Quarter<br />

BIO 12-Annual Precipitation<br />

BIO 13-Precipitation of Wettest Month<br />

BIO 14-Precipitation of Driest Month<br />

BIO 15-Precipitation Seasonality<br />

BIO 16-Precipitation of Wettest Quarter<br />

BIO 17-Precipitation of Driest Quarter<br />

BIO 18-Precipitation of Warmest Quarter<br />

BIO 19-Precipitation of Coldest Quarter<br />

DEM-Digital Elevation Model 1km<br />

Slope Angle-Derived from DEM<br />

Slope Aspect-Derived from DEM<br />

Geology-Landform Description<br />

Alternative Suitability Models<br />

Three final models were built for each species: 1)<br />

full model, 2) pruned model, and 3) topo model. The<br />

full model contained all variables. The pruned model<br />

was based on findings from a jackknife analysis, used to<br />

evaluate individual variable importance in model development.<br />

The jackknife method evaluates variable predictive<br />

strength by excluding each variable and creating<br />

a tentative model with the remaining variables (Phillips,<br />

Anderson and Schapire 2006). Then tentative models<br />

are created using each variable in isolation (Phillips,<br />

Anderson and Schapire 2006). Tentative models are<br />

then compared to the full model. The pruned model is<br />

then produced with only important predictive variables<br />

found during the jackknife analysis. The goal of the<br />

pruned model was to remove redundant variables and<br />

provide a better fit to the most important environmental<br />

predictors, when compared with the full model. The<br />

topo model used only elevation, slope angle and slope<br />

aspect as variables. There is a known correlation between<br />

elevation and climate (precipitation and temperature).<br />

The topo model was used to see how predictions<br />

of current species distributions based on topography<br />

alone (elevation, slope and aspect) compared to models<br />

with climate data (temperature, precipitation, slope and<br />

aspect). All three models are evaluated for accuracy<br />

with an AUC score.<br />

Spatial Comparison of Model Output<br />

The habitat suitability maps produced by the three<br />

models were then compared spatially to identify places<br />

of predictive agreement between models (consistently<br />

predicted present or absent), and places where predicted<br />

area of suitable habitat were in disagreement. A spatially-explicit<br />

(by pixel) comparison was performed following<br />

the methods proposed by Parolo and others<br />

(2008), which produced two output maps. The first map<br />

identifies areas of maximum agreement between the<br />

models and the second map identifies areas of minimum<br />

agreement between models.<br />

RESULTS<br />

Two hundred and eleven endemic plants were identified<br />

to have 10 or more representative location points<br />

for accurate modeling. The least number of points used<br />

for training data were seven with three test points, and<br />

the largest was 168 training points with 71 testing<br />

points. Model accuracy varied across species with AUC<br />

values ranging from 0.6423 to 1. The jackknife analysis<br />

demonstrated that slope aspect was the least predictive<br />

variable for the most species and Precipitation Seasonality<br />

(BIO15) was the highest predictive variable for the<br />

most species (Table 2).<br />

Example of ENM for Sclerocactus mesae-verdae<br />

One example of a rare endemic plant modeled is the<br />

Mesa Verde cactus (Sclerocactus mesae-verdae). Mesa<br />

Verde cactus is listed as Threatened by the US Fish and<br />

Wildlife Service and a recovery plan was written in<br />

1984 (Heil 1984). The recovery plan suggested that<br />

populations should be monitored to determine their stability<br />

(Ladyman 2004). This cactus is restricted to populations<br />

in the Four Corners region of New Mexico and<br />

Colorado. Mesa Verde cactus occurs in salt-desert scrub<br />

communities, typically in the Fruitland and Mancos<br />

shale formations, but has also been found to grow in the<br />

Menefee Formation overlaying Mancos shale (Roth<br />

2001). It is most frequently found on the tops of hills or<br />

benches and along slopes and at elevation ranging from<br />

4900 to 5500 ft (Roth 2001). Annual precipitation varies<br />

from approximately 8 to 20 cm. Average temperatures<br />

65


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 2. Total number of training and test points used in the model and the overall AUC value for the<br />

model. Predictor variable with the highest AUC value when used in isolation and the variable with the least<br />

predictive power for each species.<br />

Species<br />

# Training<br />

Samples<br />

# Test<br />

Samples<br />

Test<br />

AUC<br />

66<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Abronia argillosa 28 12 0.9759 Slope Aspect 0.5681 BIO 9 0.9039<br />

Agave utahensis ssp.<br />

kaibabensis<br />

AUC<br />

12 5 0.9933 BIO 19 0.372 BIO 15 0.9724<br />

Aliciella haydenii 14 5 0.9903 BIO 5 0.4052 BIO 11 0.8754<br />

Amsonia jonesii 47 20 0.9789 Slope Angle 0.6246 BIO 18 0.8829<br />

Amsonia peeblesii 31 12 0.9979 Slope Aspect 0.7798 BIO 14 0.9554<br />

Aquilegia grahamii 8 3 0.9955 BIO 17 0.3579 BIO 15 0.9371<br />

Aquilegia loriae 10 4 0.9994 BIO 10 0.4245 BIO 11 0.9758<br />

Argemone arizonica 8 3 0.9817 BIO 19 0.3105 Slope Angle 0.8387<br />

Argemone corymbosa ssp.<br />

arenicola<br />

7 2 0.9998 BIO 2 0.3846 BIO 16 0.9845<br />

Asclepias cutleri 20 8 0.8857 Geology 0.5897 BIO 14 0.9104<br />

Asclepias welshii 15 6 0.9802 BIO 2 0.649 BIO 11 0.9563<br />

Astragalus ampullarius 26 10 0.9957 Slope Angle 0.5299 BIO 15 0.9392<br />

Astragalus beathii 21 9 0.9994 Slope Angle 0.5269 Geology 0.9521<br />

Astragalus consobrinus 11 4 0.997 Slope Aspect 0.5 BIO 11 0.9555<br />

Astragalus cronquistii 48 20 0.9989 Slope Angle 0.6139 BIO 15 0.9437<br />

Astragalus debequaeus 63 26 0.89 Slope Aspect 0.6442 BIO 9 0.975<br />

Astragalus desperatus var.<br />

conspectus<br />

Astragalus desperatus var.<br />

petrophilus<br />

17 6 0.9989 Slope Aspect 0.4179 BIO 11 0.9945<br />

21 8 0.9991 Slope Angle 0.3093 BIO 12 0.9643<br />

Astragalus deterior 83 35 0.9997 Slope Aspect 0.6219 BIO 8 0.9842<br />

Astragalus detritalis 24 10 0.9075 Slope Aspect 0.4536 BIO 11 0.9654<br />

Astragalus duchesnensis 56 23 0.9981 Slope Angle 0.7779 BIO 7 0.9799<br />

Astragalus eastwoodiae 14 6 0.9851 BIO 10 0.4144 BIO 11 0.906<br />

Astragalus episcopus var.<br />

lancearius<br />

8 3 0.9595 BIO 8 0.4212 BIO 16 0.836<br />

Astragalus hamiltonii 11 4 0.759 Slope Angle 0.4996 BIO 16 0.9884<br />

Astragalus henrimontanensis<br />

10 4 0.9996 Slope Aspect 0.4454 Geology 0.9316<br />

Astragalus humillimus 19 7 0.9988 Slope Aspect 0.6221 BIO 4 0.9775<br />

Astragalus iodopetalus 33 14 0.9985 Slope Aspect 0.7224 BIO 7 0.9746<br />

Astragalus iselyi 16 6 0.999 Slope Aspect 0.7312 BIO 8 0.9718


Species<br />

# Training<br />

Samples<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Astragalus linifolius 14 6 0.9721 BIO 2 0.5 BIO 8 0.9492<br />

Astragalus malacoides 16 6 0.9887 Slope Aspect 0.4433 BIO 16 0.9338<br />

Astragalus micromerius 10 3 0.9897 BIO 5 0.5042 BIO 11 0.941<br />

Astragalus moencoppensis 52 22 0.9854 Slope Aspect 0.6095 BIO 16 0.9076<br />

Astragalus musiniensis 63 26 0.9868 Slope Aspect 0.4348 BIO 16 0.9137<br />

Astragalus naturitensis 62 26 0.9953 Slope Aspect 0.5935 BIO 8 0.9406<br />

Astragalus nutriosensis 21 8 0.9982 Slope Aspect 0.4817 Geology 0.9908<br />

Astragalus oocalycis 8 3 0.9958 BIO 19 0.4252 BIO 16 0.9727<br />

Astragalus perianus 20 8 0.9749 BIO 2 0.6125 Geology 0.9398<br />

Astragalus piscator 16 6 0.9999 Slope Angle 0.6152 BIO 9 0.9976<br />

Astragalus proximus 12 5 0.9772 BIO 12 0.4203 BIO 16 0.9341<br />

Astragalus rafaelensis 83 35 0.9999 Slope Angle 0.707 BIO 7 0.9709<br />

Astragalus rusbyi 21 9 0.9982 Slope Angle 0.7211 BIO 3 0.9328<br />

Astragalus schmolliae 18 7 0.9996 Slope Angle 0.7367 BIO 11 0.9848<br />

Astragalus serpens 19 7 0.9976 Slope Aspect 0.4439 BIO 8 0.9671<br />

Astragalus sesquiflorus 24 10 0.9983 Slope Aspect 0.491 BIO 6 0.9368<br />

Astragalus sophoroides 28 11 0.9994 BIO 2 0.8478 BIO 12 0.9723<br />

Astragalus striatiflorus 24 10 0.9959 Slope Aspect 0.5023 BIO 15 0.952<br />

Astragalus tortipes 8 3 0.988 Slope Angle 0.4142 Geology 0.9211<br />

Astragalus troglodytus 52 22 0.9989 Slope Angle 0.6131 BIO 3 0.9536<br />

Astragalus welshii 16 6 0.7897 Slope Aspect 0.4063 BIO 15 0.8887<br />

Astragalus wetherillii 12 5 0.9845 BIO 19 0.4486 BIO 15 0.9848<br />

Astragalus xiphoides 45 18 0.9912 Slope Angle 0.5483 BIO 1 0.9824<br />

Camissonia atwoodii 50 21 1 Slope Aspect 0.6725 BIO 12 0.9868<br />

Camissonia eastwoodiae 19 7 0.9848 BIO 17 0.5237 BIO 13 0.8891<br />

Camissonia exilis 38 16 0.9739 Slope Aspect 0.6027 BIO 15 0.948<br />

Carex curatorum 19 7 0.9907 Slope Aspect 0.2923 BIO 15 0.8895<br />

Carex specuicola 77 32 0.9969 Slope Aspect 0.4858 BIO 11 0.9395<br />

Castilleja aquariensis 168 71 0.9999 Slope Aspect 0.6332 Geology 0.9901<br />

Castilleja kaibabensis 24 9 0.9999 Slope Aspect 0.6471 BIO 15 0.9874<br />

Castilleja revealii 14 6 0.9782 Slope Aspect 0.5 Geology 0.9151<br />

Chrysothamnus molestus 35 15 0.9993 Slope Aspect 0.6306 Geology 0.9756<br />

AUC<br />

67


Species<br />

Chrysothamnus viscidiflorus<br />

ssp. planifolius<br />

# Training<br />

Samples<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

AUC<br />

8 3 0.9916 BIO 2 0.4925 Slope Angle 0.7825<br />

Cirsium chellyense 11 4 0.9864 BIO 10 0.4346 BIO 11 0.9344<br />

Cirsium murdockii 11 4 0.7394 BIO 4 0.2853 BIO 15 0.9444<br />

Cirsium perplexans 50 21 0.9924 Slope Angle 0.6892 Geology 0.9253<br />

Cirsium rydbergii 26 11 0.985 Slope Aspect 0.459 Geology 0.9099<br />

Clematis hirsutissima 19 7 0.9999 Slope Aspect 0.4112 BIO 12 0.9771<br />

Cleomella palmeriana 17 6 0.9784 Slope Angle 0.4802 BIO 13 0.889<br />

Crataegus saligna 19 8 0.9261 Slope Aspect 0.4154 BIO 6 0.9537<br />

Cryptantha atwoodii 22 9 0.9994 Slope Aspect 0.5488 BIO 16 0.9697<br />

Cryptantha capitata 33 14 0.9858 Slope Aspect 0.5139 BIO 15 0.9494<br />

Cryptantha cinerea var.<br />

arenicola<br />

12 4 0.9967 Slope Angle 0.4673 BIO 15 0.9804<br />

Cryptantha creutzfeldtii 33 14 0.9997 Slope Aspect 0.6437 BIO 6 0.9685<br />

Cryptantha elata 10 4 0.9997 BIO 1 0.402 Slope Angle 0.9762<br />

Cryptantha johnstonii 12 4 0.9987 Slope Aspect 0.2564 BIO 6 0.9714<br />

Cryptantha jonesiana 20 8 0.6423 Slope Aspect 0.2748 BIO 9 0.9782<br />

Cryptantha longiflora 17 7 0.8946 BIO 5 0.4916 BIO 15 0.8834<br />

Cryptantha mensana 9 3 0.8785 Slope Angle 0.217 BIO 15 0.9045<br />

Cryptantha osterhoutii 37 15 0.9886 Slope Aspect 0.5552 BIO 13 0.914<br />

Cryptantha paradoxa 11 4 0.9657 BIO 10 0.4083 BIO 7 0.8628<br />

Cryptantha semiglabra 21 9 1 Slope Angle 0.7076 BIO 7 0.9914<br />

Cycladenia humilis var.<br />

jonesii<br />

26 10 0.9978 Slope Aspect 0.491 BIO 13 0.9634<br />

Cymopterus duchesnensis 39 16 0.9988 Slope Aspect 0.7142 BIO 11 0.9706<br />

Cymopterus megacephalus 24 9 0.9492 Slope Aspect 0.5658 BIO 15 0.9087<br />

Cymopterus minimus 34 14 0.9916 Slope Aspect 0.4755 BIO 15 0.933<br />

Dalea flavescens 33 13 0.9489 Slope Aspect 0.5197 BIO 13 0.8887<br />

Draba graminea 10 4 0.9983 Slope Aspect 0.4768 BIO 10 0.9948<br />

Eremocrinum albomarginatum<br />

10 4 0.9826 BIO 9 0.2196 BIO 7 0.9466<br />

Ericameria zionis 8 3 0.7485 BIO 19 0.3877 Slope Aspect 0.839<br />

Erigeron kachinensis 60 25 0.9972 Slope Aspect 0.7543 BIO 15 0.9619<br />

Erigeron maguirei 21 8 0.9926 Slope Aspect 0.397 BIO 9 0.9714<br />

68


Species<br />

# Training<br />

Samples<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Erigeron mancus 17 6 0.9998 Slope Aspect 0.6755 BIO 15 0.9914<br />

Erigeron proselyticus 10 4 0.9798 BIO 12 0.3529 BIO 10 0.918<br />

Erigeron religiosus 13 5 0.9252 BIO 5 0.3833 BIO 15 0.9528<br />

Erigeron rhizomatus 23 9 0.9999 Slope Aspect 0.6843 BIO 9 0.9796<br />

Erigeron sionis 7 3 0.7639 BIO 19 0.445 BIO 15 0.8189<br />

Erigeron sivinskii 14 6 0.9732 Slope Angle 0.4754 BIO 8 0.9373<br />

Eriogonum aretioides 14 5 0.9821 Slope Angle 0.5789 BIO 15 0.9153<br />

Eriogonum bicolor 19 8 0.995 BIO 2 0.5918 BIO 4 0.9464<br />

Eriogonum clavellatum 45 19 0.9977 Slope Angle 0.6982 BIO 7 0.9668<br />

Eriogonum contortum 14 5 0.9996 BIO 2 0.6408 BIO 13 0.9855<br />

Eriogonum jonesii 26 11 0.9904 Slope Angle 0.5547 Geology 0.9317<br />

Eriogonum leptocladon<br />

var. ramosissimum<br />

AUC<br />

23 9 0.9909 Slope Aspect 0.5519 BIO 1 0.9259<br />

Eriogonum pelinophilum 111 47 0.9999 Slope Angle 0.716 BIO 7 0.9776<br />

Eriogonum ripleyi 31 13 0.9998 Slope Aspect 0.5704 BIO 15 0.9741<br />

Eriogonum scabrellum 12 5 0.9896 BIO 9 0.4757 Geology 0.9042<br />

Eriogonum subreniforme 21 9 0.8827 Slope Aspect 0.4721 BIO 12 0.9149<br />

Eriogonum tumulosum 14 5 0.9914 BIO 5 0.4753 BIO 7 0.9641<br />

Errazurizia rotundata 19 7 0.9991 BIO 3 0.7049 BIO 7 0.9724<br />

Euphorbia aaron-rossii 39 16 0.9991 Slope Aspect 0.5234 BIO 12 0.9537<br />

Euphorbia nephradenia 10 3 0.9928 Slope Aspect 0.4562 BIO 16 0.9519<br />

Gaillardia flava 14 6 0.9975 Slope Angle 0.6717 BIO 6 0.9812<br />

Gilia caespitosa 24 9 1 Slope Aspect 0.817 BIO 18 0.9578<br />

Gilia stenothyrsa 13 5 0.9989 BIO 10 0.496 BIO 9 0.9748<br />

Gilia tenuis 8 3 0.9998 Slope Aspect 0.1882 BIO 12 0.9445<br />

Glaucocarpum suffrutescens<br />

46 19 1 Slope Aspect 0.7096 BIO 9 0.9904<br />

Grindelia fastigiata 10 4 0.9902 Slope Angle 0.2632 BIO 9 0.9077<br />

Grindelia laciniata 12 5 0.9722 BIO 3 0.3623 BIO 17 0.8674<br />

Hackelia gracilenta 21 9 0.9998 Slope Aspect 0.6853 BIO 15 0.9844<br />

Hedeoma diffusa 53 22 0.9998 Slope Aspect 0.6175 BIO 15 0.9649<br />

Hedysarum occidentale<br />

var. canone<br />

21 9 0.9987 Slope Aspect 0.6461 BIO 15 0.9575<br />

69


Species<br />

# Training<br />

Samples<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Hesperodoria salicina 28 11 0.9129 Slope Aspect 0.5241 BIO 15 0.9367<br />

Heterotheca jonesii 12 5 0.9941 BIO 2 0.4925 Geology 0.8956<br />

Hymenoxys jamesii 25 10 0.9953 Slope Aspect 0.5684 BIO 11 0.9456<br />

Ipomopsis polyantha 17 6 0.9462 BIO 1 0.4734 BIO 7 0.8384<br />

Lepidium huberi 10 3 0.9728 BIO 17 0.498 BIO 9 0.9058<br />

Lepidium montanum var.<br />

neeseae<br />

AUC<br />

21 9 0.9541 Slope Aspect 0.7261 BIO 13 0.9437<br />

Lesquerella congesta 23 9 1 Slope Aspect 0.6799 BIO 6 0.9971<br />

Lesquerella kaibabensis 10 3 0.8568 Slope Aspect 0.5 Geology 0.9772<br />

Lesquerella navajoensis 12 5 0.9999 Slope Angle 0.5 Geology 0.989<br />

Lesquerella parviflora 86 36 0.9981 Slope Aspect 0.6236 BIO 15 0.9626<br />

Lesquerella pruinosa 13 5 1 Slope Aspect 0.58 BIO 16 0.9931<br />

Lesquerella vicina 24 10 0.9987 Slope Angle 0.6923 BIO 9 0.9722<br />

Lomatium concinnum 129 54 0.9998 Slope Aspect 0.6419 BIO 4 0.9586<br />

Lomatium latilobum 9 3 0.9962 BIO 17 0.4158 BIO 4 0.9019<br />

Lupinus crassus 38 16 0.9863 Slope Aspect 0.5558 BIO 8 0.9477<br />

Lygodesmia doloresensis 32 13 0.9999 Slope Aspect 0.7192 BIO 8 0.9717<br />

Mentzelia marginata 14 5 0.942 Slope Angle 0.3475 BIO 15 0.8468<br />

Mentzelia rhizomata 54 22 0.9999 Slope Aspect 0.716 BIO 15 0.9939<br />

Mentzelia shultziorum 8 3 0.6819 BIO 17 0.4103 BIO 4 0.9093<br />

Mimulus eastwoodiae 41 17 0.979 BIO 5 0.5615 Geology 0.8184<br />

Myosurus nitidus 12 4 0.79 BIO 5 0.435 Geology 0.9316<br />

Nama retrorsum 39 16 0.9892 Slope Aspect 0.5523 BIO 11 0.9098<br />

Oenothera acutissima 57 24 0.9982 Slope Aspect 0.611 BIO 13 0.9686<br />

Opuntia aurea 35 15 0.9957 Slope Aspect 0.5969 BIO 1 0.9615<br />

Oreoxis trotteri 9 3 0.997 BIO 17 0.4995 Slope Angle 0.9762<br />

Packera franciscana 10 4 0.9977 Slope Aspect 0.5 BIO 15 0.9707<br />

Parthenium ligulatum 13 5 0.9637 BIO 8 0.43 BIO 15 0.8352<br />

Pediocactus bradyi 29 12 0.9992 Slope Angle 0.7323 BIO 16 0.9645<br />

Pediocactus despainii 17 7 0.9966 Slope Aspect 0.4568 BIO 8 0.9807<br />

Pediocactus paradinei 28 12 0.9966 Slope Angle 0.7985 BIO 15 0.9701<br />

Pediocactus peeblesianus<br />

var. fickeiseniae<br />

40 17 0.9937 BIO 2 0.6728 BIO 11 0.9523<br />

70


Species<br />

# Training<br />

Samples<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Pediocactus winkleri 10 4 0.9997 Slope Aspect 0.3578 BIO 9 0.9799<br />

Pediomelum pariense 21 8 0.9839 Slope Aspect 0.4554 BIO 15 0.9679<br />

Penstemon ammophilus 19 7 0.9999 Slope Aspect 0.6623 BIO 15 0.9808<br />

Penstemon atwoodii 15 6 0.9978 Slope Aspect 0.4472 BIO 15 0.9662<br />

Penstemon bracteatus 18 7 0.9893 Slope Aspect 0.7072 BIO 9 0.9267<br />

Penstemon breviculus 9 3 0.6775 Slope Angle 0.2816 BIO 15 0.854<br />

Penstemon clutei 56 24 0.9973 Slope Aspect 0.5949 BIO 15 0.9707<br />

Penstemon distans 16 6 1 BIO 5 0.6628 BIO 1 0.9889<br />

Penstemon flowersii 19 7 1 Slope Aspect 0.7017 BIO 7 0.9976<br />

Penstemon gibbensii 8 3 1 Slope Angle 0.3782 Geology 0.9964<br />

Penstemon goodrichii 17 7 0.9999 Slope Angle 0.762 BIO 16 0.9904<br />

Penstemon grahamii 90 38 0.9996 Slope Aspect 0.69 BIO 9 0.9865<br />

Penstemon lentus var.<br />

albiflorus<br />

AUC<br />

16 6 0.9986 BIO 5 0.625 BIO 15 0.9771<br />

Penstemon marcusii 12 5 0.9991 Slope Angle 0.4632 BIO 18 0.9852<br />

Penstemon nudiflorus 79 33 0.9971 Slope Aspect 0.6484 BIO 15 0.9493<br />

Penstemon pseudoputus 39 16 0.9896 Slope Aspect 0.5151 BIO 15 0.9375<br />

Penstemon strictiformis 9 3 0.7698 Slope Aspect 0.5716 Slope Angle 0.6493<br />

Penstemon uintahensis 13 5 0.808 Slope Aspect 0.4688 BIO 15 0.9571<br />

Perityle specuicola 12 5 0.9926 Slope Angle 0.4812 BIO 8 0.9507<br />

Phacelia cephalotes 31 13 0.9851 Slope Angle 0.6395 BIO 4 0.9068<br />

Phacelia constancei 35 15 0.991 Slope Aspect 0.6325 Geology 0.952<br />

Phacelia crenulata var.<br />

angustifolia<br />

26 10 0.9787 BIO 19 0.5829 BIO 15 0.871<br />

Phacelia glechomifolia 46 19 0.9974 Slope Aspect 0.5353 BIO 15 0.9332<br />

Phacelia rafaelensis 17 7 0.9748 Slope Angle 0.5166 BIO 15 0.9301<br />

Phacelia splendens 17 6 0.9513 BIO 3 0.5761 BIO 8 0.9122<br />

Phacelia welshii 18 7 0.9987 Slope Angle 0.724 Geology 0.9767<br />

Phlox caryophylla 11 4 1 Slope Angle 0.4067 BIO 16 0.9953<br />

Phlox cluteana 23 9 0.9987 Slope Aspect 0.4581 Geology 0.9465<br />

Physaria obcordata 28 12 0.9895 Slope Aspect 0.5847 BIO 15 0.9721<br />

Physaria repanda 11 4 0.9962 Slope Angle 0.5 BIO 15 0.9436<br />

Platanthera zothecina 47 20 0.9555 Slope Aspect 0.6004 BIO 15 0.8757<br />

71


Species<br />

# Training<br />

Samples<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

# Test<br />

Samples<br />

Table 2. Continued<br />

Test<br />

AUC<br />

Least Predictive<br />

Variable<br />

AUC<br />

Highest Predictive<br />

Variable<br />

Potentilla angelliae 13 5 0.8715 Slope Angle 0.4887 BIO 9 0.9699<br />

Primula specuicola 27 11 0.9925 BIO 3 0.6973 BIO 16 0.8986<br />

Psoralidium junceum 7 3 0.9942 BIO 11 0.5 BIO 12 0.9351<br />

Psorothamnus arborescens<br />

var. pubescens<br />

Psorothamnus thompsoniae<br />

var. whitingii<br />

AUC<br />

10 4 0.9997 BIO 2 0.4595 BIO 16 0.9812<br />

12 4 0.9996 Slope Angle 0.5391 BIO 14 0.984<br />

Rosa stellata ssp. abyssa 15 6 0.999 Slope Aspect 0.5631 Geology 0.9772<br />

Salix arizonica 129 54 0.9961 Slope Aspect 0.7927 Geology 0.9772<br />

Schoenocrambe argillacea 57 24 0.9999 Slope Aspect 0.7184 BIO 9 0.9905<br />

Sclerocactus brevispinus 45 19 1 Geology*** 0.9011 BIO 7 0.9995<br />

Sclerocactus glaucus 10 4 0.9039 BIO 10 0.3434 BIO 15 0.9249<br />

Sclerocactus mesae-verdae 146 62 0.9997 Slope Aspect 0.648 BIO 6 0.9794<br />

Sclerocactus parviflorus<br />

var. intermedius<br />

11 4 0.9914 BIO 3 0.446 BIO 18 0.9473<br />

Sclerocactus sileri 29 12 0.9949 Slope Aspect 0.6299 BIO 15 0.9548<br />

Sclerocactus whipplei 35 14 0.9668 Slope Aspect 0.4761 BIO 15 0.852<br />

Sclerocactus wrightiae 81 34 0.9869 Slope Angle 0.5928 BIO 12 0.9858<br />

Shepherdia rotundifolia 91 38 0.9945 Slope Aspect 0.5764 BIO 15 0.924<br />

Silene petersonii 20 8 0.9766 Slope Aspect 0.5782 BIO 6 0.9093<br />

Silene rectiramea 11 4 1 Slope Aspect 0.5592 BIO 15 0.9866<br />

Sphaeralcea janeae 18 7 0.9984 Slope Angle 0.7861 BIO 13 0.9927<br />

Sphaeralcea psoraloides 40 17 0.9993 Slope Angle 0.5282 BIO 12 0.9812<br />

Talinum thompsonii 10 4 1 Slope Aspect 0.5 BIO 6 0.9978<br />

Thelypodiopsis juniperorum<br />

33 13 0.9989 Slope Aspect 0.6393 BIO 7 0.9586<br />

Townsendia aprica 52 21 0.9934 Slope Aspect 0.6216 BIO 11 0.9343<br />

Townsendia glabella 77 33 0.9767 Slope Angle 0.5305 BIO 15 0.9328<br />

Townsendia rothrockii 36 15 0.9902 BIO 2 0.6204 BIO 15 0.925<br />

Trifolium neurophyllum 20 8 0.9862 Slope Aspect 0.4605 BIO 3 0.9729<br />

Vanclevea stylosa 31 12 0.9924 Slope Aspect 0.6921 BIO 12 0.9618<br />

Xylorhiza glabriuscula var.<br />

linearifolia<br />

7 3 0.9981 BIO 11 0.4973 BIO 4 0.9204<br />

Zigadenus vaginatus 16 6 0.995 Slope Angle 0.5341 Geology 0.9368<br />

72


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

in the town of Shiprock near plant locations range from<br />

a high/low in January (coldest month) of 44.4 o F/17 o F to<br />

a high/low in July (hottest month) of 95 o F/58 o F (Ladyman<br />

2004). The federal listing of Mesa Verde cactus<br />

provides detailed information on habitat, soil and climatic<br />

requirements of the plant, providing great detail to<br />

compare to model data.<br />

Mesa Verde Cactus has 208 species location points.<br />

Models were developed using 146 for model training<br />

and 62 for model test validation points. The full, pruned<br />

and topo models were used to analyze suitable habitat<br />

for Mesa Verde cactus. The AUC score of the full<br />

model, incorporating all of the environmental variables,<br />

was 0.9977. The AUC score of the pruned model, developed<br />

from jackknife analysis, was 0.997. The pruned<br />

model showed geology as the highest predictive vari-<br />

able. In comparison, the topo model AUC score was<br />

0.831. The accuracy of the three niche models measured<br />

by AUC scores, demonstrated that all three models performed<br />

better than random (AUC 0.5). The full and<br />

pruned models were nearly identical in AUC scores<br />

(0.9977, 0.997). The topo model was out performed by<br />

both the full and pruned models.<br />

Variable isolation during jackknife analysis indicated<br />

geology, minimum temperature of coldest month, mean<br />

temperature of coldest quarter and temperature seasonality<br />

were the most predictive variables, with an AUC of<br />

0.9 or above (Table 3). Variable inclusion during jackknife<br />

analysis did not show any significant results; the<br />

AUC scores only changed 0.001 between models. The<br />

least predictive variable was slope aspect.<br />

Table 3: Mesa Verde Cactus model AUC scores.<br />

Sclerocactus mesae-verdae Jackknife Variable Exclusion Jackknife Variable Isolation<br />

#Training samples 146 AUC without Geology 0.997 AUC with only Slope Aspect 0.5824<br />

Iterations 500 AUC without BIO 4 0.9974 AUC with only BIO 2 0.6194<br />

Training AUC 0.9977 AUC without BIO 1 0.9975 AUC with only Slope Angle 0.7073<br />

#Test samples 62 AUC without BIO 10 0.9975 AUC with only BIO 12 0.7626<br />

Test AUC 0.9977 AUC without BIO 11 0.9975 AUC with only BIO 19 0.7773<br />

AUC Standard Deviation 0.0005 AUC without BIO 13 0.9975 AUC with only BIO 17 0.7802<br />

AUC without BIO 16 0.9975 AUC with only BIO 14 0.7907<br />

AUC without BIO 17 0.9975 AUC with only BIO 5 0.8222<br />

AUC without BIO 18 0.9975 AUC with only BIO 18 0.834<br />

AUC without BIO 19 0.9975 AUC with only BIO 13 0.8354<br />

AUC without BIO 8 0.9975 AUC with only BIO 16 0.8423<br />

AUC without BIO 9 0.9975 AUC with only BIO 7 0.8472<br />

AUC without BIO 12 0.9976 AUC with only BIO 3 0.8556<br />

AUC without BIO 14 0.9976 AUC with only BIO 10 0.8608<br />

AUC without BIO 15 0.9976 AUC with only BIO 1 0.8774<br />

AUC without BIO 2 0.9976 AUC with only BIO 9 0.8813<br />

AUC without BIO 3 0.9976 AUC with only BIO 8 0.8919<br />

AUC without BIO 5 0.9976 AUC with only BIO 15 0.8961<br />

AUC without BIO 6 0.9976 AUC with only BIO 4 0.9141<br />

AUC without BIO 7 0.9976 AUC with only BIO 11 0.9169<br />

AUC without Slope Angle 0.9976 AUC with only BIO 6 0.9655<br />

AUC without Slope Aspect 0.9976 AUC with only Geology 0.9888<br />

73


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

The habitat suitability maps also varied between the<br />

models (Figure 2). The full model (Figure 2a) and the<br />

pruned model (Figure 2b) clearly show a smaller range<br />

than the topo model (Figure 2c). The area of suitable<br />

habitat for the full model was 919,973 acres, the pruned<br />

model 1,386,261 acres and the topo model 47,110,640<br />

acres.<br />

The spatial comparison of models found 571,307<br />

acres of suitable habitat agreement between the models<br />

(Figure 2d). Disagreement analysis of the three models<br />

shows 47,081,482 acres (Figure 2e) and suggests that<br />

the topo model generated the most disagreement between<br />

models.<br />

DISCUSSION<br />

Alternative Suitability Models<br />

The use of alternative models can offer a better understanding<br />

of how environmental variables play a role<br />

in a species distribution. The comparison of the three<br />

models may provide a deeper knowledge of speciesenvironment<br />

relationships and lead to a more rigorous<br />

assessment of potential distributions. The alternative<br />

models may also identify needs in data to further understand<br />

species-environment relationships.<br />

The three models used in this analysis all performed<br />

better than random. The full and pruned models performed<br />

the best (highest AUC scores) while the topo<br />

model over predicted suitable habitat. The jackknife<br />

analysis provided a way to understand variable importance<br />

for the use in the pruned model. This may help<br />

eliminate redundant variables and an overly large<br />

model.<br />

Most pruned models showed a larger area of suitable<br />

habitat when compared with the full models. This suggests<br />

two possible interpretations: 1) the full model may<br />

be over fit, and not predicting all potential suitable habitat,<br />

or 2) the pruned model may be over predicting. Another<br />

interesting aspect of using alternative models was<br />

the use of only topographic variables in the topo model.<br />

Although this model over predicted suitable habitat for<br />

Mesa Verde cactus, other species models show a closer<br />

fit. This model may provide a way to analyze a species<br />

distribution when only elevation data are available, as<br />

Parolo and others (2008) found with Arnica montana.<br />

Spatial Comparison<br />

The spatial comparison of habitat suitability map<br />

output from the three models provides quantification of<br />

uncertainty in the model predictions of suitable habitat.<br />

Levels of uncertainty have important management implications.<br />

Areas with high model agreement that a species<br />

is present but without known occurrences of that<br />

species are target areas for field surveys. Areas of disagreement<br />

may provide insight into variables that con-<br />

tribute to uncertainty that could be better resolved<br />

through additional field work.<br />

Final Habitat Maps and Use<br />

The goal of this project was to identify potential suitable<br />

habitat for species by using alternative models<br />

evaluated with AUC scores and a spatial comparison.<br />

When comparing the three models, the AUC scores<br />

range from good to excellent for Mesa Verde cactus.<br />

The spatial comparison identified the full and pruned<br />

models as having similar predicted areas, while the area<br />

predicted by the topo model was larger. This technique<br />

provides a more rigorous analysis of the potential distribution<br />

of suitable habitat.<br />

The methods described above provide a more<br />

straightforward ecological interpretation of how the environment<br />

affects a species’ distribution. This modeling<br />

technique can also provide a better understanding of<br />

endemic plant distributions. These models can be used<br />

for future field investigations to find new populations<br />

and to identify relationships between climate, geology<br />

and topography with endemics.<br />

ACKNOWLEDGEMENTS<br />

This project was supported by the Gloria Barron Wilderness<br />

<strong>Society</strong> Scholarship and NSF grant 0753163.<br />

Comments from Dr. Deana Pennington and Daniela<br />

Roth greatly improved the manuscript and many thanks<br />

to those who have shared their data and time to improve<br />

this project.<br />

LITERATURE CITED<br />

Araujo, M. and A. Guisan. 2006. Five (or so) challenges<br />

for species distribution modeling. Journal of Biogeography<br />

33: 1677-1688.<br />

Elith, J., C. Graham, R. Anderson, M. Dudik, S. Ferrier,<br />

A. Guisan, R. Hijmans, F. Huettmann, J. Leathwick,<br />

A. Lehmann, J. Lucia, L. Lohmann, B. Loiselle,<br />

G. Manion, C. Moritz, M. Nakamura, Y. Nakazawa, J.<br />

McC Overton, A.T. Peterson, S. Phillips, K. Richardson,<br />

R. Scachetti-Pereira, R. Schapire, J. Soberon, S. Williams,<br />

M. Wisz, and N. Zimmermann, 2006. Novel<br />

methods improve prediction of species’ distributions<br />

from occurrence data. Ecography 29: 129-151.<br />

Hanley, J.A. and B.J. McNeil. 1982. The meaning<br />

and use of the area under a Receiver Operating Characteristic<br />

(ROC) curve. Radiology 29: 773-785.<br />

Heil, K.D. 1984. Mesa Verde cactus (Sclerocactus<br />

mesae-verdae) Recovery plan. U.S. Fish and Wildlife<br />

Service, Region 2, Albuquerque, New Mexico.<br />

Kartesz, J. and A. Farstad. 1999. Multi-scale analysis<br />

of endemism of vascular plant species. In: Terrestrial<br />

ecoregions of North America: A conservation assessment.<br />

Eds. Ricketts, T.H., Dinerstein, E., Olson, D.M.<br />

and C. Loucks. Island Press, Washington, D.C.<br />

74


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 2. Mesa Verde Cactus distribution maps. Full model with all variables (a), pruned model (b), topo model (c),<br />

Suitable Habitat maps Maximum Agreement (d), Suitable Habitat maps Minimum Agreement (e).<br />

75


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Ladyman, J.A.R. 2004. Status Assessment Report for<br />

Sclerocactus mesae-verdae. Prepared for The Navajo<br />

Nation Natural Heritage Program, Window Rock, AZ<br />

by JnJ Associates, LLC. Centennial, CO.<br />

Parolo, G., G. Rossi, and A. Ferrarini. 2008. Toward<br />

improved species niche modeling: Arnica montana in<br />

the Alps as a case study. Journal of Applied Ecology 45:<br />

1410-1418.<br />

Pearson, R.G., C.J. Raxworthy, M. Nakamura, and<br />

A.T. Peterson. 2007. Predicting species distributions<br />

from small numbers of occurrence records: a test case<br />

using cryptic geckos in Madagascar. Journal of Biogeography<br />

34: 102-117.<br />

Peterson, A. T., J. Soberon and V. Samcjez-Cordero.<br />

1999. Conservatism of Ecological Niches in Evolutionary<br />

Time. Science 285(5431): 1265.<br />

Peterson, A. T. and D. A. Vieglais 2001. Predicting<br />

Species Invasions Using Ecological Niche Modeling:<br />

New Approaches from Bioinformatics Attack a Pressing<br />

Problem. BioScience 51(5): 363-371.<br />

Peterson, A. T., M. Ortega-Huerta, V. Sanchez-<br />

Cordero, J. Soberon, R. Buddemeier and D. Stockwell.<br />

2002. Future Projections for Mexican Faunas Under<br />

Global Climate Change Scenarios. Nature (416).<br />

Phillips, S.J., R.P. Anderson, and R.E. Schapire.<br />

2006. Maximum entropy modeling of species geographic<br />

distributions. Ecological Modelling 190: 231-<br />

259.<br />

Ricketts, T.H., E. Dinerstein, D.M. Olson, and C.<br />

Loucks. 1999. Terrestrial ecoregions of North America:<br />

A conservation assessment. Island Press, Washington,<br />

D.C.<br />

Roth, D. 2001. Species account for Sclerocactus mesae-verdae.<br />

Navajo Nation Natural Heritage Program.<br />

Window Rock, AZ<br />

Stockwell, D. and D. Peters. 1999. The GARP Modelling<br />

System: Problems and Solutions to Automated<br />

Spatial Prediction. Intl. Journal of Geographical Information<br />

Science 13.<br />

Stockwell, D. and A.T. Peterson. 2002. Effects of<br />

sample size on accuracy of species distribution models.<br />

Ecological Modelling 148: 1-13.<br />

Welsh, S.L. 1978. Problems in plant endemism on<br />

the Colorado Plateau. Great Basin Naturalist Memoirs<br />

2:191-195.<br />

76


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Biogeography of Rare <strong>Plant</strong>s of the<br />

Ash Meadows National Wildlife Refuge, Nevada<br />

Leanna Spjut Ballard,<br />

Ballard Ecological Consulting, Millville, UT<br />

Abstract. The Ash Meadows National Wildlife Refuge encompasses more than 23,000 acres of unique desert that<br />

provides habitat for at least 25 plant and wildlife species found nowhere else in the world. Distinctive hydrology sustains<br />

high concentrations of endemic plants on the Refuge. Spring complexes, alkaline desert uplands, and velvet ash,<br />

emergent marsh and wet meadow communities known only to exist within the Refuge provide structure and habitat<br />

for several rare, endemic and endangered plants. Two studies designed to assist the Refuge with large-scale habitat<br />

restoration plans are underway. These studies include mapping all vegetation communities to a fine scale and locating<br />

and mapping the distribution of rare and listed plants on the Refuge. Mapping and classifying the vegetation communities<br />

to the alliance and association scale throughout the entire Refuge will provide a baseline of existing ecological<br />

conditions for monitoring change in the future. The rare plant surveys will also serve as a tool for monitoring the rare<br />

plants and the habitats in which they occur. Initially, vegetation classification standards were based upon community<br />

types derived from multiple earlier published sources. Due to the unique habitats and patterns of co-dominance of<br />

species occurring at the Refuge, several new alliance and association classification descriptions are being written to<br />

accurately describe plant communities.<br />

Description of the Ash Meadows National Wildlife<br />

Refuge<br />

The Ash Meadows National Wildlife Refuge<br />

(AMNWR) encompasses more than 23,000 acres of<br />

unique desert that provides habitat for at least 25 species<br />

found nowhere else in the world. The refuge may support<br />

the largest concentration of endemic species of any<br />

terrestrial landscape in the 48 contiguous United States.<br />

The Devil’s Hole pupfish (Cyprinodon diabolis) and<br />

one plant species, Ash Meadows niterwort (Nitrophila<br />

mohavenis), are listed as federally endangered, and most<br />

of the other endemic species are either listed as threatened<br />

or are managed as sensitive species. Unfortunately,<br />

past human activities in AMNWR have resulted in the<br />

invasion of numerous non-native plant species, several<br />

of which negatively affect native populations and habitats<br />

(Otis Bay and Steven Ecological Consulting 2006).<br />

Ash Meadows is essentially a watered island amidst the<br />

vast Mojave Desert. Groundwater discharging from a<br />

regional carbonate aquifer feeds the numerous springs<br />

that exist at Ash Meadows. Spring discharge maintains<br />

soil moisture in the lowlands while uplands receive water<br />

only from rainfall that averages less than 2.75 inches<br />

annually. Annual evaporation exceeds 98.50 inches.<br />

Ash Meadows is situated at approximately 2,200 feet<br />

elevation in the Mohave Desert, 40 miles east of Death<br />

Valley National Monument headquarters at Furnace<br />

Creek, California, and 90 miles northwest of Las Vegas,<br />

Nevada (Figure 1). It is located in the east-central portion<br />

of the Amargosa Desert. A series of Cambrian<br />

limestone and dolomite ridges form the eastern boundary<br />

of the Refuge near the center of the Amargosa Val-<br />

ley. The northern boundary crosses Quaternary playa<br />

deposits while the southern boundary primarily crosses<br />

Quaternary alluvial fan deposits. The pattern of sand<br />

dunes, badlands, alluvial fans, and broad meadows observed<br />

throughout Ash Meadows is the result of a history<br />

of playa deposition followed by erosion, formation<br />

of alluvial fans adjacent to surrounding ranges, and the<br />

transport of sand by both wind and infrequent flows in<br />

larger washes and drainages.<br />

The distinctive hydrology of Ash Meadows is the<br />

result of an extensive groundwater system and surface<br />

water drainage that culminates in the Carson Slough, a<br />

tributary to the Amargosa River (Otis Bay and Stevens<br />

Ecological Consulting 2006). Carson Slough is the primary<br />

drainage in Ash Meadows and is generally considered<br />

the core of the Ash Meadows ecosystem. The Crystal<br />

Spring drainage and the Jackrabbit/Big Spring drainages<br />

are significant tributaries of Carson Slough and<br />

drain large portions of the Refuge. Two primary aquifers,<br />

a regional carbonate aquifer and a local valley-fill<br />

aquifer, are present. Water is generally retained within<br />

the wetlands and alkali flats that sustain many of the<br />

Refuge’s endemic plants, as well as some endemic fish<br />

and wildlife species (BLM 2007). The persistence of<br />

this water since the late Pliocene/early Pleistocene has<br />

allowed for the continued existence of relict plants and<br />

animals which gained access to the region during pluvial<br />

climates. The isolation of these species in this harsh<br />

environment permitted their differentiation from related<br />

taxa and resulted in the distinctive character of many<br />

present-day occupants (Reveal 1980). The onset of<br />

more xeric conditions isolated Ash Meadows, thereby<br />

77


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 1. Location of the Ash Meadows National Wildlife Refuge.<br />

78


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

prohibiting genetic exchange with nearby populations,<br />

leading to progressive differentiation in plant and animal<br />

species now endemic to the area (Reveal 1980).<br />

Reveal (1980) concluded that four of the endemic Ash<br />

Meadows species (Astragalus phoenix, Mentzelia leucophylla,<br />

Grindelia fraxinopratensis and Centarium<br />

namophilum) are most closely related to congeners presently<br />

found in montane portions of the Intermountain<br />

Region. Their persistence to the present day is attributed<br />

to successful adaptation to a more xeric environment,<br />

the local persistence of water, and to relatively cool<br />

temperatures created by cool air drainage from the surrounding<br />

mountains (Beatley 1977, Reveal 1980).<br />

Current-day vegetation at AMNWR is composed of a<br />

typical Mohave creosote shrub vegetation community in<br />

addition to emergent marshes, wet meadows, distinctive<br />

spring complexes, alkaline desert uplands, and velvet<br />

ash community assemblages, several of which are<br />

known to exist only within the Refuge (Figure 2). These<br />

vegetation communities provide habitat for several rare<br />

and endangered plants, including endemic species, as<br />

well as federally-listed fish and wildlife species (Bio-<br />

West 2007).<br />

CREATION OF THE REFUGE<br />

Several legal and management documents led to the<br />

establishment of AMNWR in 1984. Devil’s Hole National<br />

Monument was declared by presidential proclamation<br />

in 1953, and federal water rights for it were adjudicated<br />

by the Supreme Court in 1976. The 1966 National<br />

Wildlife Refuge System Administration Act provided<br />

direction on Refuge management responsibilities<br />

and guidance. The Endangered Species Act of 1973, as<br />

amended, provided authority for appropriate protection<br />

and management of federally listed species. The U.S.<br />

Fish and Wildlife Service (USFWS) prepared a Warm<br />

Springs Pupfish Recovery Plan in 1976, and a Devil’s<br />

Hole Pupfish Recovery Plan in 1980. The Refuge was<br />

established on 18 June, 1984 with the purchase of<br />

12,654 acres of land from The Nature Conservancy.<br />

The Refuge now occupies a total of 23,488 acres in<br />

the Ash Meadows valley. Since designation, several<br />

documents have guided Refuge management. The 1987<br />

Ash Meadows Refuge Management Plan outlined general<br />

principles for management of Refuge ecosystems<br />

and listed species. The 1990 AMNWR Recovery Plan<br />

for the Listed Species of Ash Meadows outlined recovery<br />

needs for 12 listed species, and identified tasks to be<br />

completed to recover and downlist or delist endangered<br />

species (Sada 1990). In addition to the individual threatened<br />

and endangered plants and animals of Ash Meadows,<br />

the plan recognized the need for the recovery of<br />

Ash Meadows habitats, processes and ecosystems. The<br />

plan also included specific guidance on management<br />

objectives (Sada 1990). In 2000, an Environmental Assessment<br />

was completed (Otis Bay and Stevens Ecological<br />

Consulting 2006).<br />

ECOSYSTEM RESTORATION<br />

Managers at AMNWR are seeking to restore impacted<br />

wetland and desert upland habitats to conditions<br />

that existed 100 years ago in an effort to promote endemic<br />

species recovery. Large-scale ecosystem restoration<br />

plans include impacted habitats in the Carson<br />

Slough and Crystal Reservoir areas. Successful restoration<br />

projects have been completed at Kings Pool in the<br />

Point of Rocks area, and at Jackrabbit Spring, where<br />

visitors can view desert pupfish (Bio-West 2008).<br />

While the Refuge still supports a complex system of<br />

important native communities, portions have been significantly<br />

impacted by historic agricultural and mining<br />

activities. Impacts have included peat mining in marshlands<br />

surrounding Carson Slough in the 1960s and alfalfa<br />

farming and cattle grazing in the 1970s. These activities<br />

reduced discharge from all springs, and many<br />

spring outflows were channelized (Sada 1984). A complex<br />

irrigation system was constructed to support farming<br />

efforts, and resulting agricultural impacts included<br />

grading, cutting irrigation trenches, pumping spring<br />

pools, and creating water holding areas such as Crystal<br />

Reservoir (Otis Bay and Stevens Ecological Consulting<br />

2006). Land disturbances created by farming and grazing<br />

activities, as well as severe alteration of an ecosystem’s<br />

hydrology, can cause considerable change in<br />

vegetation community composition and allow for the<br />

encroachment of weedy and non-native species (Fraser<br />

and Martinez 2002).<br />

Two of the most obvious chronic threats to AMNWR<br />

species and ecosystems involve flow modification and<br />

land conversion associated with former agricultural development<br />

and invasive species. The most severe longterm<br />

threat to the Refuge is potential future groundwater<br />

extraction from the regional carbonate aquifer (Otis Bay<br />

and Stevens Ecological Consulting 2006). Currently, the<br />

Refuge is developing large-scale habitat restoration<br />

plans for Carson Slough, Crystal Reservoir, and other<br />

springs and areas around the Refuge. The proposed<br />

plans consider the removal of the remaining irrigation<br />

system and water control structures in an effort to restore<br />

the hydrology and geomorphology of the Refuge<br />

to a natural system. The hydrologic restoration will also<br />

support vegetation community and wildlife habitat restoration<br />

attempts. However, restoring an existing water<br />

system that has supported an area for decades could<br />

have unintended consequences.<br />

Within the Refuge, impacts resulting from the historic<br />

alterations to the landscape are evident both in the<br />

extensive monocultures of salt cedar (Tamarix ramosis-<br />

79


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 2. Vegetation community type map of the Ash Meadows National Wildlife Refuge.<br />

80


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

sima) and Russian knapweed (Acroptilon repens) found<br />

throughout Carson Slough, as well as in the parceled<br />

wet meadows created adjacent to Rogers Spring where<br />

spring water once flowed unobstructed into the slough.<br />

The Refuge is in the habitat restoration stage and will<br />

remain so for many years. Goals of the restoration plan<br />

include restoring natural hydrology and native vegetation<br />

communities, establishing a baseline of existing<br />

vegetation communities, and managing and recovering<br />

rare and endangered species occurring on the Refuge<br />

(McKelvey 2007). The U.S. Fish and Wildlife Service<br />

is conducting long-range, strategic management and<br />

restoration planning at AMNWR to accomplish the recovery<br />

goals of ecosystem and species restoration. The<br />

recovery objective for the Refuge is: “delisting for all<br />

species but the Devil’s Hole pupfish, which can only be<br />

downlisted to threatened status” (Sada 1990).<br />

PRELIMINARY VEGETATION MAPPING AND<br />

CLASSIFICATION<br />

The consulting firm BIO-WEST, Inc. undertook two<br />

studies designed to assist the AMNWR with its restoration<br />

efforts. The first study involved mapping all vegetation<br />

communities to a fine scale (0.25 acres), with the<br />

objective of providing a baseline data set for evaluation<br />

of management actions and future vegetation change,<br />

while the second included a comprehensive survey of<br />

distribution and abundance for twelve rare plant species.<br />

Completion of vegetation mapping on the Refuge has<br />

resulted in 6,237 delineated polygons (Figure 2). Of the<br />

delineated polygons, 5,913 (or 95 percent) have been<br />

assigned a preliminary alliance. Alliance assignments<br />

will be referred to as preliminary until classification is<br />

finalized. These classifications are being derived from<br />

multiple sources including Alliances of the Mojave Desert<br />

(USGS 2004), the National Vegetation Classification<br />

System (Grossman et al. 1998), and community<br />

data available on NatureServe (USGS 2004).<br />

As currently assigned, the alliances comprising the<br />

highest total acreage on the Refuge are the Atriplex confertifolia<br />

Shrubland Alliance, the Larrea tridentata-<br />

Ambrosia dumosa Shrubland Alliance, and the Isocoma<br />

acradenia Shrubland Alliance. Several delineated vegetation<br />

communities do not correspond to any previous<br />

classifications. Often this is the result of a Refuge community<br />

that has a typical dominant species occurring<br />

with an atypical co-dominant. An example of this is the<br />

Atriplex confertifolia Shrubland Alliance compared with<br />

the Atriplex confertifolia-Suaeda moquinii Shrubland<br />

Alliance. The first is a common community throughout<br />

the desert southwest. However, communities with both<br />

Atriplex confertifolia and Sueda moquinii occurring as<br />

co-dominants have not been classified. In these cases a<br />

new alliance classification may be written to best represent<br />

the vegetation community. Once final classifica-<br />

tions have been assigned, botanical descriptions will be<br />

developed for each of the newly created alliances.<br />

Association classifications are also currently in the<br />

developmental stage. Many of the common botanical<br />

associations are applicable to communities at the Refuge.<br />

However, a significant number of the delineated<br />

polygons contain associations of plants that are not<br />

commonly recognized in current classifications. These<br />

associations are being thoroughly researched in order to<br />

identify appropriate resources for classification assignments.<br />

As with the alliance classifications, we expect to<br />

develop several new association classifications to accurately<br />

describe the communities at the Refuge (BIO-<br />

WEST 2008).<br />

The vegetation mapping effort has resulted in a<br />

clearer picture of the diverse composition of vegetation<br />

communities that exist within AMNWR (Figure 2).<br />

Common vegetation communities that have been identified<br />

include Alkali Sink (an extensive shrubland community<br />

dominated by succulent shrubs such as Mojave<br />

seablite, [Suaeda moquinii] that occur adjacent to seasonally<br />

flooded wetlands and along desert washes), as<br />

well as a variety of wetland communities. Lowland Riparian<br />

Woodlands are found in the lowest elevations of<br />

the Refuge; they support a variety of canopy species<br />

such as velvet ash (Fraxinus velutina), mesquite<br />

(Prosopis), and narrow-leaf willow (Salix exigua). One<br />

of the more unique community types is the Alkali Playa<br />

community found west of Crystal Reservoir and Lower<br />

Marsh. This community may support the largest populations<br />

of rare and endemic plants on the Refuge.<br />

The western portion of the Refuge also supports<br />

well-established populations of salt cedar, Russian<br />

knapweed, and five-hook bassia (Bassia hyssopifolia).<br />

The salt cedar communities function as riparian woodlands<br />

and in some cases may provide important wildlife<br />

habitat. Several of the abandoned agricultural fields<br />

have been infested with five-hook bassia. However,<br />

these fields currently receive enough seasonal inundation<br />

to sustain recruiting populations of native wetland<br />

vegetation.<br />

Fairly intact transitional Upland and Desert Shrubland<br />

communities are present in the central and eastern<br />

portions of the Refuge. The Alkali Shrub community<br />

type transitions from a mesic phase as the topography<br />

rises in elevation from west to east. A noticeable<br />

change in vegetation composition occurs as whiteflower<br />

rabbitbrush (Chrysothamnus albidus)-dominated communities<br />

become replaced by alkali goldenbush<br />

(Isocoma acradenia) in the higher elevations where the<br />

water table is less accessible. Moving up into the alluvial<br />

fans, what was once classified as part of Creosote<br />

Shrubland is now classified as Salt Desert Shrubland<br />

composed of desert holly (Atriplex hymenelytra), shad-<br />

81


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

scale saltbush (Atriplex confertifolia), and spiny hopsage<br />

(Grayia spinosa).<br />

Creosote bush-dominated communities tend to persist<br />

in the alluvial fans east of Devil’s Hole and south<br />

along the eastern Refuge boundary in dry uplands and<br />

pavement soils. Approximately 5,000 acres of Creosote<br />

bush Shrubland is found throughout the NE corner of<br />

the refuge. It is dominated by creosote bush (Larrea<br />

tridentata) and white bursage (Ambrosia dumosa) and is<br />

one of the most common vegetation types in the Mojave<br />

Desert (MacMahon 2000).<br />

The central-eastern portions of the Refuge are home<br />

to some of the most extensive old field disturbances and<br />

remnants of agricultural activities conducted prior to the<br />

Refuge’s inception. Many of the non-native old fields<br />

contain Russian knapweed (Acroptilon repens), thistle<br />

spp. (Cirsium spp.) and tamarisk (Tamarix spp.). Old<br />

field non-native vegetation is largely the result of physical<br />

manipulation of land and land cover during agricultural<br />

uses. Old agricultural fields around the dam northeast<br />

of Point of Rocks are largely covered in annual<br />

grasses and forbs but some creosote bush and mesquite<br />

are also recolonizing. The Alkali Seep and transitional<br />

shrubland communities east of the Cold Spring private<br />

property have emerged as an important location for several<br />

endemic species such as Ash Meadows sunray, Ash<br />

Meadows blazingstar, and Ash Meadows ivesia.<br />

Alkali Meadow is a community exclusive to the<br />

Amargosa Valley and Owens Valley ecosystems. The<br />

community is a low-elevation grassland, typically with<br />

moist alkaline soils. Evaporation of surface water often<br />

leaves a crumbled salt crust over the soils. Alkali<br />

Meadow is dominated by inland salt grass (Distichlis<br />

spicata) and alkali sacaton (Sporabolis airoides). Arctic<br />

rush (Juncus arcticus) and whiteflower rabbitbrush are<br />

associated species. The federally listed spring loving<br />

centaury and Ash Meadows ivesia are found in this<br />

community. Alkali meadows are indicative of shallow<br />

water tables. Alkali flats peculiar to Ash Meadows sustain<br />

the highest concentrations of the federally-listed<br />

Amargosa niterwort.<br />

Extensive Alkali Shrubland communities dominated<br />

by Atriplex species (A. lentiformis, A. canescens, and A.<br />

confertifolia) are found in areas where groundwater is<br />

approximately 2-6 meters deep. At AMNWR, Alkali<br />

Shrubland covers 5,000 acres and comprises over 20%<br />

of the area. Alkali Meadow and Alkali Shrubland vegetation<br />

are distributed in close proximity to one another.<br />

In many places, there are raised mounds where the<br />

groundwater may be slightly deeper than surrounding<br />

alkali meadows. In these places, saltgrass, alkali sacaton,<br />

and Atriplex shrub cover increases. Other shrub<br />

species include matchbrush (Gutierrezia sarothrae),<br />

alkali goldenbush, and greasewood (Sarcobatus vermiculatus).<br />

Mesquite bosque vegetation is found pre-<br />

dominantly around spring vents and outflow channels.<br />

The dominant overstory species include mesquite<br />

(Prosopis spp.), Fremont cottonwood (Populus fremontii),<br />

and velvet ash (Fraxinus velutina).<br />

Emergent vegetation is found only at the springs and<br />

along permanent lakes and reservoirs. Emergent vegetation<br />

covers about 130 acres and comprises 0.5% of the<br />

refuge. Common species include Typha spp., spikerush<br />

(Eleocharis) spp., bulrushes, and rush species. Wetlands<br />

associated with the spring complexes include both native<br />

and non-native plant associations.<br />

RARE PLANT SURVEY METHODS<br />

Prior to the establishment of the Refuge, little quantitative<br />

information was available on the life history<br />

strategies, population genetics, demography, community<br />

associations, habitat requirements or abundance of plant<br />

species endemic to Ash Meadows. Implementation of<br />

the recovery plan for the seven listed endemic plant species<br />

will be successful only when these characteristics<br />

are known and disturbed environments are appropriately<br />

managed. Five additional at-risk plant species have been<br />

identified as species of concern at Ash Meadows, bringing<br />

the total to 12 targeted species for rare plant studies<br />

(Tables 1, 2).<br />

Beginning in 2007, BIO-WEST conducted the first<br />

comprehensive inventories of rare, endemic and listed<br />

plant species that occur within Refuge boundaries. The<br />

purpose of the rare plant studies was to obtain a baseline<br />

of existing ecological conditions and rare plant distributions<br />

as a foundation for future monitoring of changes in<br />

the status of rare plant species. The information provided<br />

by these studies will assist with planning future<br />

habitat-restoration activities and can be correlated with<br />

wildlife studies to understand benefits and potential detriments<br />

resulting from these management strategies.<br />

Floristic field surveys were initiated in March 2007<br />

and continued through late October 2008. Survey methods<br />

and intensity depended on the size of the area, investigator<br />

skill, size of the target species, and topography.<br />

Prior to conducting the field surveys, the area was<br />

analyzed to determine potential habitat for each species<br />

of interest. Potential habitat types for each species were<br />

identified in part by literature, background maps provided<br />

by Refuge staff, and observations from field visits<br />

in prior years. An intuitive controlled survey (the most<br />

commonly used and efficient method of surveying for<br />

rare plant abundance), was initially employed in 2007.<br />

The data from these initial surveys were used to design<br />

a more detailed methodology for surveying the distribution<br />

each target species during the 2008 field season.<br />

Surveyors at each population documented the estimated<br />

number of individuals, plant phenology, population distribution<br />

in terms of approximate area, and associated<br />

vegetation. As a general rule plant occurrences less than<br />

82


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. Status of Endangered, Threatened, and Sensitive species occurring in Ash Meadows National<br />

Wildlife Refuge<br />

Taxon Name Agency Status NV Heritage<br />

Ranks<br />

Scientific Vernacular FWS BLM FS NV NNPS Global State<br />

Arctomecon merriamii White bearpoppy S S W 3 3<br />

Astragalus phoenix Ash Meadows milkvetch T S F T 2 2<br />

Calochortus striatus Alkali mariposa lily S S W 2 1<br />

Centaurium namophilum Spring-loving centaury T S F T 2Q 2<br />

Cordylanthus tecopensis Tecopa birds’-beak S T 2 2<br />

Enceliopsis nudicaulis var.<br />

corrugata<br />

Ash Meadows sunray T S F T 2 2<br />

Eriogonum concinnum Darin buckwheat W 2 2<br />

Grindelia fraxinopratensis Ash Meadows gumplant T S F T 2 2<br />

Ivesia kingii var. eremica Ash Meadows ivesia T S F T 1-2Q 1-2<br />

Mentzelia leucophylla<br />

Ash Meadows blazingstar<br />

T S F T 1Q 1<br />

Nitrophila mohavensis Amargosa niterwort E S F E 1 1<br />

Phacelia parishii Parish phacelia W 2-3 2-3<br />

Salvia funerea Death Valley sage W 3 1<br />

Sisyrinchium funereum<br />

Death Valley blue-eyed<br />

grass<br />

T 2-3 1<br />

Sisyrinchium halophilum Nevada blue-eyed grass 4-5 4<br />

Sisyrinchium radicatum<br />

Spiranthes infernalis<br />

St. George blue-eyed<br />

grass<br />

Ash Meadows ladiestresses<br />

W 2?Q 1-2<br />

T 1 1<br />

FWS (Endangered Species Act administered by USDI Fish and Wildlife Service): E - endangered; T - threatened<br />

BLM/FS (USDI Bureau of Land Management, USDA Forest Service): S - special status or sensitive; W - watch<br />

NV (state of Nevada): F - fully protected<br />

NNPS (Nevada <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>): W - watch list; T - threatened; E - endangered<br />

Nevada Heritage Program ranks (Global = worldwide; State = within Nevada): 1 - critically imperiled; 2 - imperiled;<br />

3 - vulnerable; Q - taxonomic status in question; ? - rank uncertain<br />

83


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

0.25 acres in size were documented as a point feature<br />

and occurrences larger than 0.25 acres were mapped as<br />

polygon features. Specific systematic methods of sampling<br />

for the 12 rare species of interest used during the<br />

2008 field season are shown in Table 2 (Ballard 2008).<br />

RARE PLANT SURVEY RESULTS<br />

Amargosa niterwort (Nitrophila mohavensis):<br />

Listed on May 20, 1985, Amargosa Niterwort is the<br />

only endangered plant at Ash Meadows and is restricted<br />

to the Amargosa River drainage (Knight and Clemmer<br />

1987). This member of the Chenopodiaceae is an extremely<br />

hardy dwarf rhizomatous perennial that is tolerant<br />

of high soil salinity and alkalinity. It occupies the<br />

most localized habitat of any plant species endemic to<br />

Ash Meadows and is often the only species present in its<br />

habitat. N. mohavensis is found in areas with heavy salt<br />

crusts created by evaporation of standing water. These<br />

sites are characterized by barren, moist alkali flats with<br />

sandy loam soils (~57% sand) encrusted with a layer of<br />

salt with a pH near 8.4. Distichlis spicata (inland saltgrass)<br />

is sometimes found either on the periphery, or<br />

occasionally intermixed within Amargosa niterwort<br />

populations (Mozingo and Williams 1980). Without ad-<br />

equate surface water, this habitat may be taken over by<br />

saltgrass. Reveal (1978a) noted that Amargosa niterwort<br />

is sensitive to disturbance and does not reinvade sites<br />

where salt crust overlying the soil has been disturbed.<br />

Additional associated species include Atriplex confertifolia<br />

(shadscale saltbush), Mojave seablite, and a more<br />

widely distributed congener Nitrophila occidentalis<br />

(boraxweed; Soil and Ecology Research Group 2004).<br />

Two other listed species, Ash Meadows ivesia (Ivesia<br />

kingii) and Tecopa birds beak (Cordylanthis tecopensis)<br />

are also found in this type of habitat.<br />

At the time of listing, the only known location for<br />

Amargosa niterwort was Tecopa, California (Otis Bay<br />

and Stevens Ecological Consulting 2006). Since that<br />

time, several populations have been documented at the<br />

Refuge and just outside its western boundary. The Nevada<br />

Natural Heritage Program (NNHP) estimated the<br />

population of this species across its entire range at<br />

13,000 individuals in 1997 (Morefield 2001). According<br />

to the recently published five-year review for the<br />

species, the Crystal Reservoir population was estimated<br />

at 10,000 ramets (above-ground stems), and the West<br />

Refuge Boundary population was estimated at approximately<br />

50 ramets (USFWS 2007b). The two popula-<br />

Table 2. Systematic surveying protocols according to species.<br />

Scientific Name Common Name Sampling Protocol<br />

Arctomecon merriamii White bearpoppy Individual count census and transect method<br />

Astragalus phoenix Ash Meadows milkvetch Transect method<br />

Calochortus striatus Alkali mariposa lily Individual count census<br />

Centaurium namophilum Spring-loving centaury Population census and negative sampling<br />

Cordylanthus tecopensis Tecopa bird’s-beak Population census<br />

Enceliopsis nudicaulis var.<br />

corrugata<br />

Ash Meadows sunray<br />

84<br />

Transect method<br />

Eriogonum concinnum Darin buckwheat Not located<br />

Grindela fraxinopratensis Ash Meadows gumplant Population census<br />

Ivesia kingii var. eremica Ash Meadows ivesia Population census<br />

Mentzelia leucophylla Ash Meadows blazingstar Individual count census and transect method<br />

Nitrophila mohavensis Amargosa niterwort Individual count census<br />

Phacelia parishii Parish phacelia Not located<br />

Salvia funerea Death Valley sage Not located<br />

Sisyrinchium spp.* Blue-eyed grass Population census<br />

Spiranthes infernalis Ash Meadows ladies-tresses Individual count census and population census<br />

*Includes Sisyrinchium funereum, S. halophilum, and S. radicatum.


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

tions mentioned in the five-year review were the only<br />

populations known to occur on the Refuge at that time.<br />

Surveys conducted during 2008 extended the<br />

boundaries of the two known populations and added<br />

new occurrences. Some new populations were surveyed<br />

in the area just north of the outflow canal of Crystal<br />

Reservoir. In addition, occurrences noted during the<br />

2007 reconnaissance that were scattered across the west<br />

shore of Lower Marsh were resurveyed as well as populations<br />

found southwest of Crystal Reservoir. These areas<br />

included the drainage from Crystal Reservoir just<br />

north of the Refuge boundary and the Big Spring and<br />

Jackrabbit Spring drainage complex toward the western<br />

Refuge boundary. Also, a portion of critical habitat located<br />

in the west corner of the Refuge directly west of<br />

the Lower Marsh access road was mapped and inventoried.<br />

This population is referred to as the “Central<br />

Carson Slough” population by USFWS (2007b). The<br />

population was mapped in its entirety, including portions<br />

that fell just outside the Refuge boundary, and the<br />

entire population estimate for this polygon was included<br />

in the 2008 total.<br />

White Bearpoppy (Arctomecon merriamii): White<br />

bearpoppy is a Mohave Desert endemic known from<br />

Clark, Lincoln, and Nye Counties in Nevada, and from<br />

the Death Valley region of California. This species occurs<br />

in salt desert shrub communities on ridges, rocky<br />

slopes, gravelly canyon washes, and old lakebeds de-<br />

rived from carbonate rock sources, often in hard clay<br />

soils or with shadscale saltbush. It is a clump-forming<br />

perennial plant with large white flowers borne individually<br />

on the tips of leafless stems. A. merriamii can be<br />

distinguished from the golden-flowered Las Vegas bearpoppy<br />

(A. californica) by its scapose stems, larger capsules,<br />

and flower color.<br />

The NNHP rare plant fact sheet states that there have<br />

been approximately 129 occurrences documented<br />

throughout its Mojave Desert range. The estimated<br />

range-wide population is > 20,000 individuals (Morefield<br />

2001) (Table 3). While there is no documented<br />

evidence of the number of known individuals within the<br />

Refuge prior to this study, it is believed that the distribution<br />

has remained limited with low abundance of individuals<br />

(H. Hundt, AMNWR, 2007, pers. comm.).<br />

Field crews conducted reconnaissance in areas of<br />

potential habitat for this species between 2,000 and<br />

6,200 feet in elevation and within Salt Desert Scrub<br />

communities on alluvial gravel substrates. These areas<br />

included the northernmost portion of the Refuge just<br />

north and south of the Invite Road, the area surrounding<br />

Devils Hole, the alluvial fans surrounding Point of<br />

Rocks, and the extreme southeast corner of the Refuge.<br />

No plants were found during these searches. However,<br />

several previously undocumented populations were discovered<br />

and surveyed throughout the 2008 field season.<br />

These included the area just south of Peterson Road<br />

Table 3. Population estimates for surveyed plant species at Ash Meadows NWR.<br />

Scientific Name Common Name Most Recent Population<br />

Estimate<br />

85<br />

2008 Survey Population<br />

Estimate AMNWR<br />

Arctomecon merriamii White bearpoppy 20,000* 193<br />

Astragalus phoenix Ash Meadows milkvetch 1800 11,643<br />

Calochortus striatus Alkali mariposa lily unknown 6984<br />

Centaurium namophilum Spring-loving centaury 4290* 4,468,571<br />

Cordylanthus tecopensis Tecopa bird’s-beak 4379* 829,918<br />

Enceliopsis nudicaulis var.<br />

corrugata<br />

Ash Meadows sunray 1849 50,954<br />

Grindela fraxinopratensis Ash Meadows gumplant 81,000 376,632<br />

Ivesia kingii var. eremica Ash Meadows ivesia 3862 486,798<br />

Mentzelia leucophylla Ash Meadows blazingstar 358 3763<br />

Nitrophila mohavensis Amargosa niterwort 10,050 78,406<br />

Sisyrinchium spp.* Blue-eyed grass unknown 99,822<br />

Spiranthes infernalis Ash Meadows ladies-tresses 1107 14,209<br />

* Range-wide estimate


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

near the Cold Spring private property, just northwest of<br />

Longstreet Road but south of Peterson Road, and near a<br />

large spring drainage between the eastern Refuge border<br />

and Longstreet Road. Many of the occurrences were<br />

found within habitat for species such as Ash Meadows<br />

sunray, Ash Meadows milkvetch, and Ash Meadows<br />

blazingstar. These habitats typically included mesic Alkali<br />

Shrublands with sandy soils and occasional deep<br />

washes. There is some evidence from the NNHP that<br />

occasionally moist sandy soils could serve as potential<br />

habitat indicators for this species (Moorefield 2001).<br />

The broad transect survey methods for the Ash Meadows<br />

sunray and Ash Meadows milkvetch likely contributed<br />

to the discovery of the new populations. The total<br />

population surveyed in 2008 was approximately 193<br />

individuals (Table 3). There is some potential for locating<br />

additional populations as the Ash Meadows sunray<br />

surveys are completed during the 2009 field season.<br />

Ash Meadows Milkvetch (Astragalus phoenix):<br />

Ash Meadows milkvetch is endemic to AMNWR. A<br />

federally protected species, it has been documented<br />

fairly extensively in the past, including studies directed<br />

at recovery of the species. The NNHP documents 13<br />

occurrences for a total estimated population of 1,943<br />

individuals within the Ash Meadows area (Morefield<br />

2001). Previous surveys conducted by the USFWS in<br />

2000 documented several populations within the Refuge<br />

totaling 1,800 individuals (Pavlik and Stanton 2006).<br />

Prior to 2006 populations were known from south of<br />

Rogers Spring and west through the northern portion of<br />

Purgatory, within the Cold Springs private property,<br />

south of Bradford Spring, east and west of Ash Meadows<br />

Road, and north and south of South Spring Meadows<br />

Road. Survey areas included potential habitat consisting<br />

of alkaline soils, desert washes, and barren flats<br />

(Reveal 1978b). Because this is commonly known to<br />

occur in habitats similar to those of Ash Meadows Sunray,<br />

both species could be surveyed together.<br />

Several new populations of Ash Meadows milkvetch<br />

were discovered during the Ash Meadows Sunray transect<br />

surveys. Large populations were discovered adjacent<br />

to the Cold Spring private property. Other notable<br />

populations were inventoried in the area between<br />

Rogers and Longstreet Springs, directly west of the<br />

junction of Ash Meadows Road and South Spring<br />

Meadows Road, and west of Jack Rabbit Spring. The<br />

estimated total population is approximately 11,643 individuals<br />

(Table 3).<br />

Alkali Mariposa Lily (Calochortus striatus):<br />

NNHP reports only four occurrences of alkali mariposa<br />

lily across its entire known range in Clark and Nye<br />

counties, Nevada and adjacent California. Morefield<br />

(2001) lists the estimated population of the species as<br />

“unknown.”<br />

Several populations documented during the 2007<br />

reconnaissance at AMNWR were surveyed in 2008.<br />

These populations included a number of locations immediately<br />

south of Collins Ranch, just west of Warm<br />

Springs and north of the access road to Bole Spring.<br />

Several new populations were located and surveyed including<br />

one at the bend in West Spring Meadows Road.<br />

A large population was surveyed within an Alkali<br />

Shrubland community east of Crystal Reservoir. An additional<br />

population was mapped and surveyed in the<br />

southeast corner of the Refuge. The recorded population<br />

for species surveyed during this study totals 6,984 individuals<br />

(Table 3).<br />

Spring-loving Centaury (Centaurium namophilum):<br />

Spring-loving centaury is an annual plant that<br />

is endemic to AMNWR and its immediate surroundings.<br />

It is currently listed as a threatened species by<br />

the USFWS. The last confirmed survey reported by<br />

NNHP was in 1986 and documented 19 occurrences for<br />

an estimated population in excess of 4,290 individuals<br />

(Morefield 2001). The draft five-year review mentions<br />

six mapped populations within the Refuge totaling over<br />

2,900 acres in comparison to the approximately 29 acres<br />

last reported to the NNHP (Morefield 2001, USFWS<br />

2008). According to the draft five-year review, population<br />

trends are insufficiently documented (USFWS<br />

2008).<br />

Survey area criteria for this species included seeps,<br />

wet meadows, and spring channel banks throughout<br />

AMNWR. In 2008, spring-loving centaury was found<br />

to be very widespread across the Refuge, populating<br />

habitats from seasonally flooded wetlands to seasonally<br />

moist Alkali Meadows and the edges of some Alkali<br />

Shrubland communities. It appeared that nearly any location<br />

on the Refuge containing surface or sub-surface<br />

water at any time during the year would produce a<br />

population. As surveys continued, a blooming trend for<br />

certain populations became apparent. Blooms were seen<br />

in “waves” for individual populations and subpopulations,<br />

or different parts of a single population would<br />

bloom at different times during the season.<br />

In the Peterson Reservoir area, extensive populations<br />

extended throughout surrounding drainages. As in the<br />

Rogers Spring and Carson Slough drainages, populations<br />

extend until they encounter what may be hydrologic<br />

barriers. Observed occurrences were so extensive<br />

that it became necessary to map areas of nonoccurrence.<br />

The total population from the 2008 surveys<br />

has been estimated at 4,468,571 individuals.<br />

Clearly the current population estimates are a significant<br />

increase from the last confirmed survey data provided<br />

to the NNHP (USGS 2004). It is clear that populations<br />

of this annual plant fluctuate widely from season<br />

to season; however, the likelihood that the number of<br />

86


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

individuals would dip as low as previously recorded<br />

estimates seems doubtful.<br />

Tecopa Bird’s-Beak (Cordylanthus tecopensis):<br />

Tecopa bird’s-beak is a hemiparasitic summer annual<br />

plant that is a Nevada Sensitive Species. It is known<br />

from approximately ten extant occurrences across a narrow<br />

range in California (Death Valley) and Nevada<br />

(AMNWR) (Morefield 2001). Its habitat includes Mohave<br />

Desert scrub and alkali flats and meadows below<br />

2,700 feet. It always grows with Distichlis spicata,<br />

which may be its principal host. Tecopa bird’s-beak is<br />

also a known associate of the spring-loving centaury<br />

and often occurs within the same habitat types (Otis Bay<br />

and Stevens Ecological Consulting 2006). Population<br />

estimates provided by the NNHP document >4,379 total<br />

individuals.<br />

Because of their similar habitat requirements, populations<br />

of Tecopa bird’s-beak were mapped and surveyed<br />

in conjunction with spring-loving centaury in<br />

2008. New occurrences were discovered along the<br />

shores of lower Crystal Marsh as well as on the west<br />

side of the marsh within old agricultural fields. The agricultural<br />

field population extended intermittently to the<br />

western Refuge boundary. In addition, a significant<br />

population was found associated with a new springloving<br />

centaury population in a narrow band of velvet<br />

ash located northwest and southeast of Collins Ranch.<br />

This area appears to be an important site for multiple<br />

rare and endemic species at the Refuge. The total population<br />

of Tecopa bird’s beak documented in the 2008<br />

field season is approximately 829,918 individuals.<br />

Ash Meadows Sunray (Enceliopsis nudicaulis var.<br />

corrugata): Ash Meadows sunray is an endemic variety<br />

of a widely distributed species that has been listed as<br />

threatened by the USFWS. The variety is almost strictly<br />

endemic to Ash Meadows with a few individuals reported<br />

from outside the Refuge in eastern California. It<br />

is largely restricted to strongly alkaline, poorly drained,<br />

saline soils associated with springs and dry washes but<br />

with the water table some distance below the surface.<br />

Lower elevation alkali clay soils in Ash Meadows have<br />

a shallow underlying water table that makes the habitat<br />

unsuitable. This species is associated with Ash Meadows<br />

milkvetch, shadscale saltbush, matchbrush, alkali<br />

goldenbush, basin yellow cryptantha (Cryptantha confertifolia)<br />

and white bearpoppy at elevations from 2,100<br />

to 2600 feet. It is generally found on dry to sometimes<br />

moist sites that are on open, hard, white clay hills with<br />

calcareous hardpans. Populations on the Refuge are<br />

found in occasionally moist alkaline soils, spring and<br />

seepage areas, and dry desert washes. The plants can<br />

also occasionally be found in salt desert shrubland and<br />

desert pavement habitats (Morefield 2001; Otis Bay and<br />

Stevens Ecological Consulting 2006). The last con-<br />

firmed population estimates for this plant were reported<br />

following its listing as a threatened species.<br />

The 2008 survey was directed at locating the plants<br />

throughout all potential habitat types occurring within<br />

the Refuge. Cruise transects 40 meters apart were used<br />

to survey large tracts of Ash Meadows. Of the more<br />

than 9,000-acres of potential habitat, nearly 6,000 acres<br />

have been surveyed to date. Ash Meadows sunray has<br />

been found throughout the areas mapped by the Refuge<br />

in 2006. In several cases known populations have been<br />

extended beyond previous distribution boundaries. New<br />

occurrences were documented west of known populations<br />

mapped along Ash Meadows Road, as well as on<br />

the alluvial fans east of Point of Rocks and south of<br />

Jackrabbit Spring. A single occurrence was also documented<br />

adjacent to Lower Crystal Marsh. The preliminary<br />

population estimate, calculated with approximately<br />

two-thirds of the survey complete, is 50,954 individuals.<br />

The remaining survey area includes habitat within the<br />

central portion of the Refuge that has long been known<br />

to support this plant. It is likely that upon completion of<br />

the surveys, the final population estimate will increase<br />

by several thousand.<br />

Ash Meadows Gumplant (Grindelia fraxinopratensis):<br />

Ash Meadows gumplant was listed as threatened<br />

by the USFWS in 1985. The plant is considered an<br />

endemic species primarily occurring within AMNWR<br />

with a limited distribution in neighboring Inyo County,<br />

California. NNHP documented 16 occurrences across<br />

the known range for the species and estimated a population<br />

of more than 13,000 individuals in 1986 (Morefield<br />

2001). The USFWS estimated the Refuge’s population<br />

at approximately 81,000 individuals following a 1998<br />

survey (USFWS 2007a).<br />

Distribution data provided by the USFWS in 2000<br />

and Refuge staff in 2006 indicate that populations of G.<br />

fraxinopratensis occur in spring drainages and marsh<br />

habitats throughout the Refuge. Notably, both data sets<br />

show a significant presence of this species along the<br />

Fairbanks and Rogers Spring drainages. However, BIO-<br />

WEST botanists involved with the 2007 reconnaissance<br />

and the 2008 surveys indicate that plants found at these<br />

locations are actually not members of this species. At<br />

this time BIO-WEST has been unable to confirm occurrences<br />

north or west of the Warm Springs complex.<br />

The Alkali Meadows south of Crystal Reservoir contain<br />

very large populations of Ash Meadows gumplant.<br />

Known populations were also documented in the Alkali<br />

Meadows of the Big Spring/Jackrabbit Spring drainage<br />

complex both east and west of South Spring Meadows<br />

Road. Another known population was inventoried south<br />

of Ash Meadows Road as it intersects South Spring<br />

Meadows Road. New populations were surveyed between<br />

the Warm Springs Complex and West Spring<br />

87


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Meadows Road, as well as just north of Devils Hole<br />

Road in the Collins Ranch area, east of Crystal Reservoir<br />

and between the south end of Lower Crystal Marsh<br />

and the Refuge boundary. Due to a short bloom time,<br />

several occurrences west of the Refuge office were not<br />

surveyed in 2008. The total estimated population at this<br />

time is 376,632 individuals, a number that will likely<br />

increase upon completion of surveys.<br />

Ash Meadows Ivesia (Ivesia kingii): Endemic to the<br />

Refuge, Ash Meadows ivesia is a federally-listed threatened<br />

species. It is a perennial plant with a prostrate<br />

growth form. As is the case with many of the endemic<br />

Refuge plants, there is little current information regarding<br />

its abundance. According to the USFWS, the distribution<br />

of this species is limited to Nye County, Nevada,<br />

and likely limited to within AMNWR boundaries. Population<br />

estimates by the NNHP in the 1980s indicated<br />

there were as many as nine occurrences; these areas<br />

contained an estimated 3,862 individuals (Morefield<br />

2001).<br />

The majority of the 2008 surveys confirmed populations<br />

occurring within historic distribution areas on the<br />

Refuge. However, field crew members were unable to<br />

locate plants within several previously documented areas<br />

believed to contain populations and extant populations<br />

appeared significantly smaller. However, a visit in<br />

the Jackrabbit Spring area led to the discovery of several<br />

new populations. New populations, fairly large in size,<br />

were also surveyed in close proximity to West Spring<br />

Meadows Road where it makes a sharp turn to the west.<br />

Also, multiple, substantial populations were located and<br />

surveyed between Crystal Reservoir and West Spring<br />

Meadows Road. Another population was documented in<br />

a seepage area interrupting an upland habitat just east of<br />

the Cold Spring private property. Finally, populations<br />

occurring just north and south of Big Springs Road were<br />

surveyed. The current estimated population of Ash<br />

Meadows ivesia is 486,798 individuals (Table 3).<br />

Ash Meadows Blazingstar (Mentzelia leucophylla):<br />

This endemic biennial herb was listed as a threatened<br />

species in 1985. The plant’s distribution appears to be<br />

strictly limited to areas within the Refuge (Morefield<br />

2001; Otis Bay and Stevens Ecological Consulting<br />

2006). Recent information regarding species abundance<br />

is limited. Surveys in 1986 documented M. leucophylla<br />

from 8 locations with an estimated population of 358<br />

individuals (Morefield 2001). Distribution maps provided<br />

by the Refuge from 2006 internal surveys show<br />

confirmed populations at Purgatory, the Warm Springs<br />

Complex, and along West Spring Meadows Road.<br />

Additional populations of Ash Meadows blazingstar<br />

were discovered during surveys for Ash Meadows sunray<br />

and Ash Meadows milkvetch. Because of its biennial<br />

life form, M. leucophhylla was observed in rosette<br />

as well as in flower. Starting in mid-March, BIOWEST<br />

field crews began recording populations south of Peterson<br />

Road near the Cold Spring private property boundary.<br />

In addition, new blazingstar populations were documented<br />

just south of Rogers Spring and west of Longstreet<br />

Road, intermixed with a large white bearpoppy<br />

population southeast of the Warm Springs Complex access<br />

road and directly north of the “T” junction of South<br />

Spring Meadows Road and West Spring Meadows<br />

Road. These populations were surveyed and their<br />

boundaries extended. The estimated population is 3,763<br />

individuals (Table 3).<br />

Blue-Eyed Grass Species (Sisyrinchium species):<br />

Previously, three species of blue-eyed grass were<br />

thought to occur at the Refuge: Death Valley blue-eyed<br />

grass (Sisyrinchium funereum), Nevada blue-eyed grass<br />

(S. halophilum), and St. George blue-eyed grass (S.<br />

radicatum) (C. Baldino 2007 pers. comm.). However,<br />

Cholewa (in letter, 2003) suggested that S. halophilum<br />

probably does not extend as far south as the refuge. She<br />

based her statement on the discovery that many herbarium<br />

specimens that had been identified as that species<br />

were incorrect. Blue-eyed grass species are extremely<br />

difficult to differentiate, so there is ongoing debate and<br />

confusion as to exactly which species actually exist on<br />

the Refuge.<br />

Such problems are not new in Sisyrinchium. Specieslevel<br />

taxonomy of Sisyrinchium has long been disputed.<br />

Recent molecular work has clarified the limits of the<br />

genus and helped identify important morphological<br />

characters delineating it from closely related genera<br />

(Karst, unpublished). However, much more phylogenetic<br />

work, based on cladistic analysis of molecular<br />

data, is needed to understand species relationships<br />

within the genus (Cholewa and Henderson 2002).<br />

Species of Sisyrinchium are not easily distinguished.<br />

White flowers may occur in otherwise blue-flowered<br />

species, and vivipary occasionally occurs, where plants<br />

produce seeds that germinate before they detach from<br />

the parent. Furthermore, vegetative characteristics,<br />

while distinctive in some species, may overlap greatly in<br />

wide-ranging species. Writers of past floras sometimes<br />

were unaware of such phenotypic plasticity, or were<br />

inconsistent in their use of terminology. Some taxonomists<br />

have thought differences too subtle and chosen to<br />

lump species (Cholewa and Henderson 1984, 2002).<br />

Because of the taxonomic confusion surrounding the<br />

blue-eyed grass plants growing within refuge boundaries,<br />

the BIO-WEST 2008 field surveys combined all<br />

occurrences of blue-eyed grass into a Sisyrinchium spp.<br />

category.<br />

As documented by Cholewa and Henderson (2002),<br />

Sisyrinchium funereum populations occur mostly within<br />

Death Valley. These populations contain numerous individuals.<br />

Sisyrinchium radicatum is more widely distributed,<br />

growing in Clark, Nye, and Lincoln Counties,<br />

88


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Nevada, and Washington County, <strong>Utah</strong>. In Clark<br />

County, it occurs in Pine Canyon and Ash Spring and is<br />

also known from Pine Creek and Red Rock Canyon-<br />

Calico Basin in the Spring Mountains. In Nye County, it<br />

is known from Big Springs in Ash Meadows, and Pahrump<br />

Valley. In Lincoln County, it is known from Pahranagat<br />

lakes and Pahranagat Valley – Ash Springs. The<br />

threats to these populations are currently unknown. Sisyrinchium<br />

demissum is a closely related species that<br />

overlaps the known range of the two species known to<br />

occur on AMNWR.<br />

Known populations of blue-eyed grass were visited<br />

and surveyed throughout the Refuge. Several previously<br />

undiscovered populations were documented including<br />

occurrences just south and east of Jackrabbit Spring.<br />

Large populations were surveyed in the area directly<br />

south of Crystal Reservoir, expanding habitat from the<br />

2007 findings. New populations were also located adjacent<br />

to the Cold Springs private property, with the large<br />

Ash Meadows ladies-tresses population northeast of<br />

Rogers Spring but south of Longstreet Road, and in a<br />

spring drainage leading from the eastern Refuge border<br />

in the north section of the Refuge. The estimated blueeyed<br />

grasses population is 99,822 individuals (Table 3).<br />

Ash Meadows ladies-tresses (Spiranthes infernalis):<br />

A Refuge endemic, Ash Meadows ladies- tresses<br />

is currently being considered for Federal listing (Otis<br />

Bay and Stevens Ecological Consulting 2006). NNHP<br />

survey records from 1998 show 15 locations with an<br />

estimated population of 1107 individuals (Morefield<br />

2001).<br />

New populations documented during the 2008 survey<br />

were found in several seep habitats The population was<br />

estimated at 14,209 individuals. Several occurrences<br />

were observed late in the field season when the plants<br />

had passed the flowering period and will be revisited<br />

during 2009. Because there can be significant variability<br />

in the number of individuals that bloom from season to<br />

season, it may be necessary to revisit surveyed locations<br />

to determine an accurate population estimate after new<br />

locations are added.<br />

The 2008 rare plant surveys of the twelve sensitive<br />

and endemic species that exist at AMNWR revealed<br />

larger populations and new locations of additional populations<br />

of these species than had been previously documented<br />

and aided in determining clearer population estimates,<br />

population location boundaries, and the associated<br />

vegetation communities these species exist in. Additional<br />

planned surveys in 2009 may aid in more accurately<br />

determining the biogeography of the rare, endemic<br />

and listed <strong>Plant</strong>s of the Ash Meadows National<br />

Wildlife Refuge.<br />

REFERENCES<br />

Ballard, L.S. 2008. Sampling protocols for rare<br />

plants at Ash Meadows National Wildlife Refuge. BIO-<br />

WEST, Inc.<br />

Beatley, J.C. 1977. Threatened plant species of the<br />

Nevada Test Site, Ash Meadows, and central-southern<br />

Nevada. 66 pps.<br />

BIO-WEST, Inc. 2007. Ash Meadows National<br />

Wildlife Refuge. 2007 Draft Progress Report. Logan,<br />

<strong>Utah</strong>.<br />

BIO-WEST, Inc. 2008. Ash Meadows National<br />

Wildlife Refuge. 2008 Draft Progress Report. Logan,<br />

<strong>Utah</strong>.<br />

[BLM] U.S. Bureau of Land Management. 2007.<br />

Amargosa River Area of Critical Environmental Concern<br />

Implementation Plan. Barstow (CA): BLM. 19 p.<br />

plus appendices and maps.<br />

Cholewa, A. F. and D.M. Henderson. 1984. Biosystematics<br />

of Sisyrinchium Section Bermudiana<br />

(Iridadeae) of the Rocky Mountains. Brittonia 36: 342-<br />

364.<br />

Cholewa, A.F. and D.M. Henderson. 2002. Sisyrinchium.<br />

Pp 351-371. In: Flora of North America Editorial<br />

Committee. Flora of North America North of<br />

Mexico. Volume 26. Magnoliophyta: Liliidae: Liliales<br />

and Orchidales.<br />

Fraser J. and C. Martinez C. 2002. Restoring a desert<br />

oasis. Endangered Species Bulletin 27:18–19.<br />

Grossman, D.H., D. Faber-Landgendoen, A.S.<br />

Weakley, M. Anderson, P. Bourgeron, R. Crawford, K.<br />

Gooding, S. Landaal, K. Metzler, K. Patterson, M. Pyne,<br />

M. Reid, and L. Sneddon. 1998. International classification<br />

of ecological communities: Terrestrial vegetation of<br />

the United States. Volume I: The national Vegetation<br />

Classification Standard. The Nature Conservancy, Arlingon,<br />

Va.<br />

Knight, T.A., and G.H. Clemmer. 1987. Status of<br />

Populations of the Endemic <strong>Plant</strong>s of Ash Meadows,<br />

Nye County, Nevada. Reno, NV: U.S. Fish and Wildlife<br />

Service, Great Basin Complex. Also available online at<br />

http://heritage.nv.gov/reports/ashmtext.pdf<br />

MacMahon, J.A. 2000. Warm Deserts. Pages 285-<br />

322 in North American Terrestrial Vegetation. M.G.<br />

Barbour and W.D. Billings eds. 2 nd edition. Cambridge<br />

University Press, Cambridge, UK.<br />

McKelvey S. 2007. The feasibility of restoring historic<br />

Carson Slough. Las Vegas: U.S. Fish and Wildlife<br />

Service. 16 p.<br />

Morefield, J.D., ed. 2001. Nevada Rare <strong>Plant</strong> Atlas.<br />

Compiled by the Nevada Natural Heritage Program.<br />

Portland, OR: Fish and Wildlife Service, U.S. Department<br />

of the Interior, Fish and Wildlife Service, Portland,<br />

Oregon and Reno, Nevada. http://heritage.nv.gov/atlas/<br />

atlastxt.pdf.<br />

89


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Mozingo, H.N. and M. Williams, 1980. Threatened<br />

and endangered plants of Nevada. An illustrated manual.<br />

U.S. Fish and Wildlife Service, Portland, OR, and<br />

U.S. Bureau of Land Management, Reno, NV.<br />

Otis Bay and Stevens Ecological Consulting. 2006.<br />

Ash Meadows geomorphic assessment and biological<br />

assessment: final report. Las Vegas: U.S. Fish and<br />

Wildlife Service. 149 p. plus appendices.<br />

Pavlik, B.M. and A.E. Stanton. 2006. Managing<br />

populations of rare plants at Ash Meadows National<br />

Wildlife Refuge. I. Demographic Survey and Habitat<br />

Quality Assessment to Recover Astragalus phoenix.<br />

Draft. San Francisco: BMP Ecosciences. 33 p. including<br />

appendices.<br />

Reveal, J.L. 1978a. Status report on Nitrophila mohavensis<br />

Munz and Roos. US Fish and Wildlife Service,<br />

Portland, OR.<br />

Reveal, J.L. 1978b. Status report on Astragalus<br />

phoenix. US Fish and Wildlife Service, Portland, OR.<br />

Reveal, J.L. 1980. Biogeography of the Intermountain<br />

Region. A speculative appraisal. Mentzelia 4: 1-92.<br />

Sada, D.W. 1990. Recovery Plan for the Endangered<br />

and Threatened Species of Ash Meadows, Nevada. U.S.<br />

Fish and Wildlife Service. Reno, Nevada.<br />

Soil Ecology and Research Group. 2004. Demographics<br />

and ecology of the Amargosa niterwort<br />

(Nitrophila mohavensis) and Ash Meadows gumplant<br />

(Grendelia fraxinopratensis) of the Carson Slough area.<br />

16 pps. Webpage: http://www.serg.sdsu.edu/SERG/<br />

restorationproj/mojave%20desert/deathvalleyfinal.htm<br />

[USFWS]. U.S. Fish and Wildlife Service. 2007a.<br />

Ash Meadows gumplant (Grindelia fraxinopratensis)<br />

five-year review: summary and evaluation. Las Vegas:<br />

Nevada Fish and Wildlife Office, Southern Nevada<br />

Field Office. Pp. 5-7.<br />

[USFWS]. U.S. Fish and Wildlife Service. 2007b.<br />

Amargosa niterwort (Nitrophila mohavensis) five-year<br />

review: summary and evaluation. Las Vegas: Nevada<br />

Fish and Wildlife Office, Southern Nevada Field Office.<br />

Pp. 5-7.<br />

[USFWS]. 2008. Spring loving centaury (Centaurium<br />

nomophilum) draft five-year review: summary and<br />

evaluation. Las Vegas: Nevada Fish and Wildlife Office,<br />

Southern Nevada field office. Pp. 6-7.<br />

[USGS] U.S. Geological Survey. 2004. Mojave Desert<br />

ecosystem program: central Mojave vegetation database.<br />

Final report. Flagstaff (AZ): Colorado Plateau<br />

Field Station, USGS Southwest Biological Science Center.<br />

251 p.<br />

90


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Assessing Vulnerability to Climate Change Among the<br />

Rarest <strong>Plant</strong>s of Nevada’s Great Basin<br />

Steve Caicco,<br />

Conservation Planner, Planning Branch, National Wildlife Refuge System,<br />

Portland, OR<br />

,<br />

Abstract. The Great Basin of Nevada provides habitat for many narrowly distributed endemic plant species. To assess<br />

the vulnerability of 33 of the rarest of these plants to climate change, I used the elevation range of all reported<br />

locations as a surrogate measure of their bioclimatic envelopes. The results show that 14 of these taxa occur on or<br />

near the valley floors, nine taxa occur in montane habitats, and 10 taxa occur at high elevations. While the majority of<br />

the 33 taxa are restricted to highly specialized edaphic habitats, valley endemics are distributed through a smaller elevation<br />

span than montane or high elevation endemics. In addition, valley habitats have less variability in slope and<br />

aspect and their highly specialized habitats do not occur above the valley floor. These habitat restrictions are likely to<br />

constrain migration in response to climate change. Montane and high elevation habitats are more diverse topographically<br />

and, although often specialized, are more common both locally and regionally. This imposes fewer constraints<br />

on natural migration and offers more conservation options in the face of climate change. Our inability to accurately<br />

predict the actual parameters of climate change and its effects at a scale relevant to rare species will require a comprehensive<br />

inventory and monitoring effort to identify those species affected by climate change. An integrated long-term<br />

seed storage program will ensure adequate representation for genetic conservation.<br />

Pollen, woodrat midden, tree-ring, and lake level<br />

data spanning the past 50,000 years has demonstrated<br />

that the Great Basin is highly sensitive to climatic<br />

change (Wharton et al. 1990). During the 20 th Century,<br />

an average annual warming of 0.3 to 0.6 °C occurred<br />

and projections for the next century are for an additional<br />

rise of 2 to 5 °C (Cubashi et al. 2001; Wagner 2003).<br />

Alterations to the precipitation regimes are harder to<br />

predict, but seem likely to include a greater proportion<br />

falling as rain, decreasing winter snowpack, and earlier<br />

arrival of spring conditions, thereby affecting runoff and<br />

plant phenology (Mote et al. 2005; Snyder and Sloan<br />

2005).<br />

The basin-and-range topography that characterizes<br />

the Great Basin of Nevada has generated much interest<br />

among biogeographers and numerous seminal works<br />

have been published focused on the distribution and relationship<br />

of its vascular flora or specific aspects of<br />

plant endemism in this region (Billings 1978; Charlet<br />

1996; Harper and Reveal 1978; Holmgren 1972a; Pavlik<br />

1989; Reveal 1979; Shreve 1942; Wells 1983). A published<br />

proceedings of a symposium on intermountain<br />

geography includes numerous papers on many aspects<br />

of biogeography in the Great Basin (Harper and Reveal<br />

1978). Several sources of information are available on<br />

the rare plants of Nevada (Morefield 2001; Mozingo<br />

and Williams 1980) or parts of Nevada (Anderson et al.<br />

1991; Spahr et al. 1991; Weixelman and Atwood undated).<br />

A conservation blueprint for the Great Basin has<br />

been prepared that includes general discussions of the<br />

ecological systems represented and their associated species<br />

conservation targets and identifies a portfolio of 20<br />

priority landscape scale conservation sites; climate<br />

change and adaptation options are also discussed but no<br />

explicit assessment of species vulnerability was conducted<br />

(Nachlinger et al. 2001).<br />

Narrowly endemic plants are expected to be at far<br />

greater risk of extinction from climate change than are<br />

more widespread plants because of their limited range,<br />

small populations, and genetic isolation (Committee on<br />

Environment and Natural Resources 2008; Peters and<br />

Darling 1985). Alpine plants are often identified as at<br />

particular risk due to isolation and lack of an “escape<br />

route” (Grabherr et al. 1995; Halloy and Mark 2003).<br />

Among the observed and predicted effects of climate<br />

change on plant species are phenological changes, trophic<br />

level disruptions, range shifts and contractions, and<br />

extinctions (Parmesan 2006). Documented effects include<br />

an accelerated trend in upward shift of alpine<br />

plants in the Swiss Alps over the last few decades<br />

(Walther et al. 2005), a significant upward shift in optimum<br />

elevation of forest plants in Europe in the 20 th century<br />

(Lenoir et al. 2008), a decline of arctic-alpine plants<br />

from 1989-2002 in Glacier National Park (Lesica and<br />

McCune 2004), and an advance in the mean flowering<br />

dates of lilac and honeysuckle in the western United<br />

States of 2 and 3.8 days per decade, respectively (Cayan<br />

et al. 2001).<br />

Extinction is predicted for 3–21 percent of the flora<br />

in Europe, 38–45 percent in the Cerrado of Brazil, 32–<br />

91


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

63 percent in the alpine flora of New Zealand, and 41–<br />

51 percent of endemics in South Africa and Namibia<br />

(Fischlin et al. 2007). In the Great Basin, range contractions<br />

or extinctions have been predicted for higher elevation<br />

mammals, butterflies, birds and plants (Murphy<br />

and Weiss 1992; Van de Ven et al. 2007; Wagner 2003).<br />

Populations of pika (Ochotona princeps) have already<br />

been reduced by 28 percent compared to the number of<br />

populations known earlier in the 20 th century (Beever et<br />

al. 2003). Grayson (2000) has posited that during the<br />

Middle Holocene (5,000–8,000 years before present), a<br />

period characterized by a decrease in summer precipitation<br />

and an increase in temperatures, the small mammal<br />

fauna of the Great Basin decreased in species richness<br />

and evenness as a result of a series of local extinctions<br />

and near-extinctions coupled with an increase in taxa<br />

well-adapted to xeric conditions.<br />

The objective of this study was to conduct an initial<br />

assessment of the vulnerability of the rarest plants of the<br />

Great Basin of Nevada to climate change based on geographical<br />

and ecological data from the literature, stored<br />

in data bases and files maintained by the Nevada Natural<br />

Heritage Program, Carson City, Nevada, and stored<br />

in files maintained by the U.S. Fish and Wildlife Service<br />

at the Nevada Fish and Wildlife Office, Reno, Nevada.<br />

STUDY AREA<br />

The study area encompassed the Great Basin within<br />

the State of Nevada. The term “Great Basin” was first<br />

used in 1844 by the explorer John Fremont in reference<br />

to the large closed hydrologic basin lying between the<br />

Sierra Nevada of California and Nevada to the west and<br />

the Wasatch Front of <strong>Utah</strong> to the east with slight extensions<br />

into Oregon and Idaho (Tingley and Pizarro 2000).<br />

This analysis focuses on the floristic Great Basin within<br />

Nevada (Holmgren 1972a), an area of roughly 54,741<br />

km 2 that includes all of Nevada north of the Mojave Desert<br />

with the exception of the Carson Range along the<br />

eastern side of Lake Tahoe (Figure 1).<br />

In general, precipitation increases and temperature<br />

decreases with elevation in the Great Basin, although<br />

physiographic factors can exacerbate temporal and spatial<br />

variation in climatic patterns. The complex terrain,<br />

with its large differences in altitude and the consequent<br />

distortion of air currents creates high variability in local<br />

precipitation and short periods of intense rainfall followed<br />

by very long periods without precipitation (Hidy<br />

and Kleiforth 1990). Rapid heat loss at night results in<br />

cool air descending to valley floors where pooling in<br />

closed basins creates diurnal temperature inversions<br />

(Beatley 1975). Along the southern boundary of the<br />

study area, air and soil temperature regimes on two valley<br />

floor sites, separated by only 90 m horizontally and<br />

1.5 m vertically, were found to influence the distribution<br />

of dominant shrub species. The composition of herb-<br />

Figure 1. Location of the study area in the Great Basin<br />

of Nevada, an area of about 54,741 km 2 ; the larger<br />

dark outline is the boundary of the Great Basin Restoration<br />

Initiative as delineated by the U.S. Bureau of<br />

Land Management based primarily on floristic similarity.<br />

aceous perennials and winter annuals on the same two<br />

sites was similar, although temperature influenced the<br />

initiation of vegetative growth in herbaceous perennials<br />

and the germination success of winter annuals (Beatley<br />

1969, 1975). Predicting local climates and the climatic<br />

responses of highly localized endemic plant populations,<br />

therefore, is at best a challenging approximation.<br />

METHODS<br />

I used ecologic and geographic data stored by the<br />

Nevada Natural Heritage Program and the U.S. Fish and<br />

Wildlife Service, including various status assessment<br />

reports prepared for many of the rare plant taxa, to determine<br />

the reported minimum and maximum elevations<br />

of all known populations of the rarest plants within the<br />

study area. All taxa ranked as G1, G1G2, or T1 were<br />

included in this study. G1 ranked species are considered<br />

critically imperiled based on a very high risk of extinction<br />

due to extreme rarity (often five or fewer populations),<br />

very steep declines, or other factors; T1 ranks are<br />

applied to infraspecific taxa that meet the same criteria<br />

as for the G1 ranks (Natureserve 2009). I assessed a total<br />

of 167 reported locations of 33 taxa with ranks of<br />

G1, G1G2, or T1 ranks (Table 1).<br />

I used the reported minimum and maximum elevation<br />

of all known locations for each of the 33 taxa as a<br />

surrogate for the combined effects of temperature and<br />

92


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

precipitation, the principal climatic factors affecting<br />

plant distribution in the Great Basin (Billings 1949;<br />

Comstock and Ehleringer 1992; Fautin 1946). Van de<br />

Ven and others (2007) used actual temperature and precipitation<br />

data to construct bioclimatic models of potential<br />

climatic change in the White Mountains of California<br />

and Nevada at the western edge of the Great Basin,<br />

but such data are relatively sparse for the interior Great<br />

Basin. Climate stations are typically clustered near<br />

towns and skewed toward lower elevations; there are<br />

few climate stations above 2500 m. Moreover, due to<br />

the complex topography of the Western United States<br />

and the coarse resolution of most climate models, even<br />

the best climate models display biases at regional scales<br />

(Bonfils et al. 2008). All 33 taxa are highly localized<br />

endemics with geographic ranges that are considerably<br />

smaller than could be predicted by the best regional climate<br />

models.<br />

While approaches that incorporate additional factors<br />

such as slope, aspect, and geologic substrate have been<br />

used with some success to develop predictive models of<br />

potential habitat for a few rare plants in the Great Basin<br />

of <strong>Utah</strong> (Aitken and others 2007), field surveys conducted<br />

by experienced botanists for many of these 33<br />

taxa have consistently found that only a small portion of<br />

their predicted range contains suitable habitat. Therefore,<br />

the objective of this study was to compare the climatic<br />

niches of rare plants, rather than to develop predictive<br />

distribution models.<br />

RESULTS<br />

The rarest plants of the Great Basin of Nevada (i.e.,<br />

those with a G1, G1G2, or T1 rank) can be placed into<br />

three broad elevation bands. Those bands reflect occurrence:<br />

1) below the lower limits of tree distribution on<br />

or near valley floors; 2) within a narrow montane zone<br />

dominated by pinyon-juniper woodlands and associated<br />

shrublands or subalpine; or 3) near or above timberline<br />

(Figure 2, Table 1).<br />

The lowest elevation group is comprised of 14 taxa<br />

that have a median population altitude below 2000 m.<br />

All but one of these taxa are endemic to Nevada (Figure<br />

2). Eleven taxa, Eriogonum tiehmii, E. argophyllum,<br />

E. ovalifolium var. williamsiae, E. diatomaceum, Castilleja<br />

salsuginosa, Johanneshowellia crateriorum,<br />

Boechera falcifructa, Mentzelia argillicola, M. tiehmii,<br />

Frasera gypsicola, and Potentilla basaltica have a reported<br />

elevational amplitude of less than 244 m, with<br />

nine of these distributed across less than 129 m of elevation<br />

(Figure 2). Two additional species, Sclerocactus<br />

blainei and Mimulus ovatus, are reported from an elevational<br />

amplitude of less than 50 m. The lone anomaly to<br />

the general pattern of restricted elevational range among<br />

the lowest elevation group was Penstemon floribundus,<br />

with populations spanning 1,009 m.<br />

The middle elevation group is comprised of nine<br />

taxa, all but one of which are thought to occur only in<br />

Nevada (Figure 2). The median altitudes of all known<br />

populations of these taxa fell between 2,000 m and<br />

2,746 m, although three taxa have some populations<br />

near or above 3,000 m. Six of these taxa, Tonestus graniticus,<br />

Lewisia maguirei, Collomia renacta, Eriogonum<br />

microthecum var. arceuthinum, E. douglasii var.<br />

elkoense, and Trifolium andinum var. podocephalum<br />

had elevational amplitudes of less than 280 m. This relatively<br />

narrow elevation span is comparable to those<br />

typical of the valley taxa, and an argument could be<br />

made to make only two groupings based on a division at<br />

2,500 m (Figure 2). Such a division would, however,<br />

mask an ecological distinction based on a notable difference<br />

between valley and montane endemics in their edaphic<br />

specialization, discussed in more detail in the following<br />

section. The remaining three taxa, Penstemon<br />

tiehmii, P. pudicus, and P. moriahensis had reported<br />

elevational amplitudes of 640 m, 782 m, and 1,128 m,<br />

respectively.<br />

The highest elevation group is comprised of ten taxa,<br />

seven of which are considered endemic to Nevada. Only<br />

three of the ten taxa, Draba serpentina, Boechera<br />

ophira, and Ipomopsis congesta var. nevadensis had<br />

reported elevational amplitudes of less than 200 m. Two<br />

taxa, Primula capillaris and Penstemon rhizomatosus,<br />

had reported elevational amplitudes of 381 m and 366<br />

m, respectively. The remaining five taxa, Viola lithion,<br />

Eriogonum holmgrenii, Potentilla cottami, Polemonium<br />

chartaceum, and Cymopterus goodrichii had reported<br />

elevational amplitudes between 747 m and 1,158 m.<br />

Only three, Polemonium chartaceum, Eriogonum holmgrenii,<br />

and Draba serpentina, were restricted to sites<br />

above treeline (Figure 2).<br />

DISCUSSION<br />

Overview<br />

Many of the rarest plant taxa in the Great Basin of<br />

Nevada are found below the lower limits of tree distribution.<br />

These low elevation taxa (median elevation below<br />

2,000 m) had the narrowest bioclimatic envelope, as<br />

estimated from their reported elevational amplitudes,<br />

with 11 of 14 (79 percent) spanning less than 244 m. If<br />

other elevation bands are considered, 17 of the 20 taxa<br />

(85 percent) with an elevational span of less than 280 m<br />

have a median elevation below 2,377 m (Figure 2).<br />

Among the ten highest elevation taxa, only three (30<br />

percent) have a reported elevational amplitude of less<br />

than 200 m (one of these is likely more widely distributed<br />

as discussed below), and four of them (40 percent)<br />

are known to occur over more than 700 m of elevation.<br />

These results oppose the prediction that alpine species<br />

are at particular risk due to isolation and lack of an<br />

93


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

“escape route” (Grabherr et al. 1995; Halloy and Mark<br />

2003). Perhaps this can be attributed to unusual features<br />

of plant endemism in the western United States in general<br />

and, more specifically, in the Intermountain West.<br />

Kruckeberg (1986) reformulated earlier works characterizing<br />

the processes of soil formation and vegetation<br />

development (Jenny 1941; Major 1951) to account for<br />

plant diversity in any region. He noted that topography,<br />

parent material, and the timing of geological processes<br />

or events created a patchiness or discontinuity of edaphic<br />

phenomena that creates additional opportunity for<br />

biological discontinuity. i.e., speciation. He termed endemic<br />

taxa that result from this process “geoedaphics”<br />

(Kruckeberg 1986). Rajakaruna (2004) provided a<br />

review that emphasized the role that unusual soil conditions<br />

play in the diversification of plant species.<br />

Although Kruckeberg (1986) emphasized the role of<br />

bedrock (and especially serpentine) outcrops in the evolution<br />

of geoedaphics, in an earlier paper Kruckeberg<br />

and Rabinowitz (1985), cast a broader net with respect<br />

to narrowly distributed endemics (sensu Mason<br />

1946a,b), noting that unique taxa associated with<br />

“gypsum, serpentine, limestone, alkaline and heavy<br />

metal soils are well known to field botanists in many<br />

parts of the world.” While few, if any, serpentine outcrops<br />

are known from the Great Basin in Nevada, the<br />

Calcareous Mountains Section of the Intermountain region,<br />

which lies primarily in the eastern half of Nevada<br />

and adjacent southwestern <strong>Utah</strong>, is recognized as the<br />

richest area of the Great Basin for plant endemism<br />

(Holmgren1972a). Elsewhere, examples of Great Basin<br />

plants restricted to exposures of carbonate bedrock are<br />

Figure 2. Elevation ranges of reported localities of the 33 rarest plants (NatureServe G1, G1G2, and T1 ranks) in the<br />

Great Basin of Nevada. The rectangles in each box-and-whisker plot show the 50 percent of the populations that occur<br />

between the 25 th and 75 th percentiles and the median elevation. Species are ranked by median elevation. Circles<br />

indicate plants known from only a single site. Shading shows plants endemic to Nevada.<br />

94


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. Rare <strong>Plant</strong> Taxa in the Great Basin of Nevada Assessed<br />

for Vulnerability to Climate Change.<br />

Taxon<br />

No. of<br />

Pop’s a<br />

95<br />

Life<br />

Form b<br />

High Elevation Endemics (n=9)<br />

Habitat(s)<br />

Polemonium chartaceum H. Mason 1 (6) pf Rocky slopes, talus, fellfields<br />

Eriogonum holmgrenii Reveal 4 pf Quartzite/limestone outcrops, slopes, ridges<br />

Draba serpentina (Tiehm & P. Holmgren) Al-<br />

Shehbaz & Windham<br />

4 pf Quartzite slopes and cliffs<br />

Penstemon rhizomatosus N.H. Holmgren 6 pf Limestone talus and cliffs<br />

Cymopterus goodrichii Welsh & Neese 7 pf Quartzite and limestone talus<br />

Boechera ophira (Rollins) Al-Shehbaz 13 pf Steep slate or limestone scree/talus<br />

Ipomopsis congesta (Hook.) V.E. Grant var.<br />

nevadensis (Tidestr.) Tiehm<br />

1 pf Carbonate-derived soils or scree?<br />

Potentilla cottamii N.H. Holmgren 2 (2) pf Quartzite cracks and crevices<br />

Primula capillaris N.H. Holmgren & A. Holmgren<br />

4 pf Moist meadows in glacial till<br />

Montane Endemics (n=10)<br />

Viola lithion N.H. Holmgren & P.K. Holmgren 5 pf Limestone/dolomite cracks, crevices,<br />

ledges<br />

Penstemon moriahensis N.H. Holmgren 8 pf Gravelly and/or silty carbonate soils<br />

Penstemon tiehmii N.H. Holmgren 3 pf Soil pockets on steep volcanic talus/scree<br />

Penstemon pudicus Reveal & Beatley 6 pf Volcanic rocky soils, crevices, boulder<br />

piles<br />

Lewisia maguirei A.H. Holmgren 8 ge Carbonate scree/shallow soils steep slopes/<br />

ridges<br />

Tonestus graniticus (Tiehm & L.M. Shultz) G.L.<br />

Nesom & D. Morgan<br />

1 pf Granite cliffs and outcrops<br />

Collomia renacta E. Joyal 2 an Volcanic lithosols<br />

Eriogonum microthecum Nutt. var. arceuthinum<br />

Reveal<br />

Trifolium andinum Nutt. var. podocephalum<br />

Barneby<br />

2 pf Tuffaceous knolls, bluffs, and rocky flats<br />

1 pf Tuffaceous bluffs and soils<br />

Eriogonum douglasii Benth. var. elkoense Reveal 1 pf Sandy to gravelly flats and slopes<br />

Valley Endemics (n=14)<br />

Eriogonum tiehmii Reveal 6 pf Rocky clays derived from mixed sedimentary<br />

rock<br />

Castilleja salsuginosa N.H. Holmgren 2 pf Seasonally moist alkaline clays/siliceous<br />

geothermal sinter<br />

Penstemon floribundus D. Danley 8 pf Volcanic talus, slopes, or colluvium<br />

Johanneshowellii crateriorum Reveal 7 an Sandy pumice flats and slopes


Taxon<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. continued<br />

No. of<br />

Pop’s a<br />

Life<br />

Form b<br />

Habitat(s)<br />

Boechera falcifructa (Rollins) Al-Shehbaz 9 pf Zonal soils with big sagebrush zone<br />

Mentzelia argillicola N.H. Holmgren & P.K.<br />

Holmgren<br />

5 pf Alkaline clays/silts of Pliocene lake beds<br />

Frasera gypsicola (Barneby) D.M. Post 10 pf Alkaline clays/silts of Pliocene lake bed;<br />

spring mounds<br />

Mentzelia tiehmii N.H. Holmgren & P.K. Holmgren<br />

7 pf Alkaline clays/silts of Pliocene lake beds<br />

Eriogonum argophyllum Reveal 1 an Siliceous geothermal sinter<br />

Sclerocactus blainei S.L. Welsh & Thorne 3 ps Alkaline volcanic and calcareous clay soils<br />

Mimulus ”ovatus” c 9 pf Sandy to gravelly flats and slopes<br />

Eriogonum ovalifolium Nutt. var. williamsiae<br />

Reveal<br />

Eriogonum diatomaceum Reveal, J. Reynolds &<br />

Picciani<br />

1 pf Siliceous geothermal sinter<br />

11 pf Diatomaceous earth deposits<br />

Potentilla basaltica Tiehm & Ertter 9 pf Moist alkaline meadows<br />

a <strong>Number</strong> of reported populations in Nevada included in study; populations outside of Nevada included are shown in<br />

parentheses.<br />

b pf = perennial forb; ge = geophyte; an = annual; ps = perennial succulent.<br />

C Recent taxonomic revisions have left this western Nevada endemic, formerly included in Mimulus ovatus, without a<br />

name.<br />

well-documented in the White Mountains of California<br />

and Nevada (Marchand 1973; Mooney 1966; Mooney et<br />

al. 1962; Morefield 1992; Wright and Mooney 1965)<br />

and on altered andesites in western Nevada (Billings<br />

1950). In an analysis of plant distributions in the Mojave-Intermountain<br />

transition zone, Meyer (1978) found<br />

that endemic plants showed a high degree of habitat specialization<br />

and that edaphically restricted species were<br />

much better represented in xeric, than in mesic, community<br />

types.<br />

Kruckeberg and Rabinowitz (1986) also provided<br />

case histories of Astragalus phoenix and Mentzelia leucophylla,<br />

both narrow edaphic endemics known from<br />

flats, washes, and knolls of calcareous alkaline soils at<br />

Ash Meadows National Wildlife Refuge in Nye County,<br />

Nevada. Ash Meadows lies on the periphery of, and<br />

shares many ecological features with, the adjacent Great<br />

Basin; paleosoils on which these two species occur are<br />

the partially dissected remnants of a large Pleistocene<br />

playa. Numerous examples also exist of endemism or<br />

rarity in Great Basin plants associated with soils derived<br />

from volcanic ash (Grimes 1984), sand dunes (Holm-<br />

gren 1979; Pavlik 1989a), geothermal features (Holmgren<br />

1972b; Reveal 1972, 1981), Pliocene and Pleistocene<br />

lake and playa sediments, including gypsum<br />

mounds, (Forbis 2007; Holmgren and Holmgren 2002;<br />

Reveal 1972), diatomaceous earth deposits (Reveal et al.<br />

2002), and pumice deposits (Reveal 2004a). A detailed<br />

discussion of examples of edaphic endemism among the<br />

rarest plants of the valley, montane, and high elevation<br />

habitats of the study area is presented in subsequent sections.<br />

In most cases, the specialized habitats of these taxa<br />

are more properly characterized as a substrate rather<br />

than a well-developed soil. Sand deposits, shallow<br />

gravel sinters, pumice fields, and volcanic ash exposures<br />

in the valleys and the scree and talus slopes of the<br />

mountains are typically dynamic, unstable substrates<br />

shaped by erosional processes. Mineral material dominates<br />

the soil profile, little organic material is present,<br />

and soil horizons are poorly differentiated, if they are<br />

even present. In a few cases, such as paleosoils developed<br />

on ancient lake beds, moist alkaline clays, or playa<br />

edges, the soils are better developed, although in the<br />

96


case of the former lake beds a duripan is common within<br />

a short distance of the surface and the surface itself<br />

may be armored with desert pavement. The marked<br />

aridity of the Great Basin also slows the rate of soil development.<br />

While the degree of soil development has a<br />

substantial influence on the floristic composition and<br />

structure of more common, widespread plant communities,<br />

it appears to be less important in specialized edaphic<br />

endemics. In these habitats, physical soil factors<br />

(or perhaps, in some cases, soil chemistry) may have<br />

greater influence on the ability of species to establish<br />

and persist.<br />

Valley Endemics<br />

The 14 rarest plants below 2000 m all occur on valley<br />

floors within a matrix of zonal vegetation dominated<br />

by Artemisia tridentata ssp. wyomingensis or various<br />

salt desert shrubs (Figure 2, Table 1). Most occur on<br />

azonal soils developed from surficial deposits that overlie<br />

the underlying bedrock, in some cases by thousands<br />

of meters (Table 1). Eriogonum diatomaceum, for example,<br />

is restricted to diatomaceous earth deposits<br />

(Reveal et al. 2002). Mentzelia argillicola, M. tiehmii,<br />

and Frasera gypsicola are typically found on sediments<br />

comprised of calcareous silts, clays, and air-deposited<br />

ash beds that accumulated in Middle Pliocene-Early<br />

Pleistocene lakes (Tschanz and Pampeyan 1970), although<br />

the latter two species also occur on Pliocene<br />

spring mounds with high gypsum content (Forbis 2007).<br />

Frasera gypsicola is also rarely found in saline bottomlands<br />

(Smith 1994). Johanneshowellia crateriorum is<br />

known only from sandy pumice flats and slopes (Reveal<br />

2004a) associated with the Quaternary Lunar Crater volcanic<br />

field (Kleinhampl and Ziony 1985). The habitats<br />

of Sclerocactus blainei and Mimulus ovatus have been<br />

described as igneous or calcareous gravels with a clay<br />

matrix (Heil and Porter 1994; Welsh and Thorne 1985)<br />

and sandy to gravelly slopes derived from siliceous<br />

sinter or hydrothermally-altered andesite (Morefield<br />

2001), respectively. Eriogonum argophyllum, E. ovalifolium<br />

var. williamsiae, and Castilleja salsuginosa are<br />

all associated with geothermal features, either growing<br />

in siliceous sinter gravels (Erigonum spp.), or in moist<br />

alkaline clays or weathered travertine (Castilleja)<br />

(Holmgren 1972b; Reveal 1972, 1981). Potentilla basaltica<br />

is also restricted to alkaline wet meadows<br />

(Tiehm and Ertter 1984).<br />

The remaining low elevation species do not fit the<br />

same pattern of adaptation to azonal soils in the valleys.<br />

Boechera falcifructa appears to be the only lower elevation<br />

species that occurs on zonal soils; the association of<br />

known populations with cryptogammic soils crusts<br />

within the Artemisia tridentata ssp. wyomingensis zone<br />

suggests that it may once have been more widely distributed<br />

but subsequently reduced by livestock tramp-<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

97<br />

ling (Morefield 1997). The only plant among the lowest<br />

elevation group to grow only on soils directly weathered<br />

from bedrock is Eriogonum tiehmii, which occurs on<br />

rocky clay soils derived from interbedded sedimentary<br />

rocks, including claystones, shales, tuffaceous sandstones<br />

and limestones (Morefield 1995; Reveal 1985).<br />

Penstemon floribundus is unique among the lower elevation<br />

taxa for the breadth of its altitudinal distribution<br />

(Figure 2); it is reported to occur on a wide variety of<br />

substrates derived from volcanic rocks (Danley 1985;<br />

Knight 1988). Extensive inventories for P. floribundus,<br />

endemic to the remote Jackson Range in northwestern<br />

Nevada, have not been conducted and the species may<br />

be more common*.<br />

Montane Endemics<br />

The nine plant species in this group occur within the<br />

narrow montane zone dominated by various species of<br />

Artemisia or the extensive woodlands of Pinus monophllya<br />

and Juniperus osteosperma typical of mountain<br />

ranges in the central Great Basin (Figure 2, Table 1). In<br />

higher mountains, these forests may be comprised of<br />

other conifers, including Abies concolor, P. flexilis, and<br />

less commonly, P. longaeva (Charlet 1996). In contrast<br />

to the valley endemics, eight of the nine montane taxa<br />

occur either on poorly developed soils directly weathered<br />

from underlying bedrock, or in scree, talus, or bedrock<br />

ledges, cliffs, and crevices (Table 1). The exception<br />

is Eriogonum douglasii var. elkoense reported from<br />

sandy to gravelly flats and slopes with Artemisia nova<br />

and mixed grasses (Reveal 2004b). The primary substrate<br />

affinities of the other eight taxa include tuffaceous<br />

volcanic sediments (Trifolium andinum var. podocephalum<br />

(Barneby 1989) and Eriogonum microthecum var.<br />

arceuthinum (Reveal 2004b)), volcanics (Collomia renacta<br />

(Joyal 1986), Penstemon pudicus (Reveal and<br />

Beatley 1971), and P. tiehmii (Holmgren 1998)), granite<br />

(Tonestus graniticus (Tiehm and Shultz 1985)), and carbonates<br />

(P. moriahensis (Holmgren 1979), Lewisia<br />

maguirei (Holmgren 1954; Williams 1981), and Viola<br />

lithion (Holmgren 1992)). In general, these substrates<br />

are common regionally and locally and the presumed<br />

rarity of these taxa is most likely determined by other<br />

ecological or historical factors.<br />

High Elevation Endemics<br />

Five of the nine high elevation endemics show a<br />

preference toward a particular geologic substrate (Table<br />

1). Those reported to occur only on quartzite or other<br />

siliceous substrates include Draba serpentina (Al-<br />

*Surveys conducted subsequent to the preparation of this<br />

manuscript have confirmed P. floribundus to be more common<br />

and to occur on the highest peaks of the Jackson Range<br />

(A. Tiehm, personal communication).


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Shehbaz and Windham 2007; Tiehm and Holmgren<br />

1991), Boechera ophira (Morefield 2003) and Potentilla<br />

cottamii (Holmgren 1987). Penstemon rhizomatosus is<br />

reported only from carbonate substrates (Holmgren<br />

1998) while Cymopterus goodrichii is known from slate<br />

and limestone sedimentary rocks (Welsh and Neese<br />

1980). Ipomopsis congesta var. nevadensis may also<br />

occur on carbonate substrates but no systematic surveys<br />

have been completed (Morefield 2001). Eriogonum<br />

holmgrenii is reported from quartzitic, carbonate, and<br />

granitic substrates (Goodrich 1979; Reveal 1965), while<br />

Polemonium chartaceum occurs on rhyolite in the<br />

Sweetwater Range of California (Hunter and Johnson<br />

1983), on open slopes of metavolcanics and nonbasic<br />

rocks at the summit of White Mountain Peak in the<br />

White Mountains of California (Crowder and Sheridan<br />

1972; Morefield 1992; Rundel et al. 2008; Van de Ven<br />

et al. 2007), and on granitic rocks in the Boundary Peak<br />

area of the White Mountains in California and Nevada<br />

(Kartesz 1987; Pritchett and Paterson 1998). Similar to<br />

their montane counterparts, these taxa are typically<br />

found on poorly developed soils derived directly from<br />

weathered bedrock or in association with scree, talus, or<br />

bedrock ledges, cliffs, and crevices. These substrates<br />

and habitats are common regionally and locally and the<br />

presumed rarity of these taxa is most likely determined<br />

by other ecological or historical factors. The sole exception<br />

to this among the high elevation species is Primula<br />

capillaris, which occurs on moist meadow soils derived<br />

from glacial till (Holland 1995; Holmgren and Holmgren<br />

1974).<br />

Vulnerability to Climate Change<br />

The vulnerability of any plant species or population<br />

to climate change is influenced by many factors in addition<br />

to climate and substrate, including physiological<br />

tolerances, life-history strategies, phenological plasticity,<br />

relative probabilities of extinctions and colonizations,<br />

dispersal abilities, and disruptions in plantpollinator<br />

or herbivorous insect-host plant interactions<br />

(Parmesan 2006). Unfortunately, few data exist on most<br />

of these factors for any plants, including the 33 plants<br />

examined herein. This analysis, therefore, provides preliminary<br />

predictions that will be tested as climate<br />

change progresses over the coming decades.<br />

Previous studies have concluded that with progressive<br />

climate change we can expect ongoing upward<br />

shifts in both forest and alpine plant species and subsequent<br />

declines in arctic-alpine plants at the southern<br />

margin of their range (Lenoir et al. 2008; Lesica and<br />

McCune 2004; Walther et al. 2005). The results reported<br />

here suggest that losses of endemic plants in the<br />

Great Basin may be highest within the lower elevation<br />

sagebrush and salt desert zones. This is consistent with<br />

similar results reported from <strong>Utah</strong> where the highest<br />

levels of plant endemism were found between 1000 m<br />

and 2000 m (Ramsey and Shultz 2004). Meyer (1978)<br />

also found that the percentage of edaphically restricted<br />

species in the Mojave-Intermountain transition zone<br />

dropped sharply with an increase in altitude. This may<br />

be because xeric environments tend to be more heterogeneous<br />

than mesic habitats in response to variables<br />

other than moisture and heterogeneous environments<br />

tend to restrict the migration of specialized plant species,<br />

thus reinforcing endemism.<br />

Previously published studies of climate change in the<br />

Great Basin have largely focused on the montane biogeography<br />

of birds and mammals (e.g., Beever 2003;<br />

Brown 1971, 1978; Lawlor 1998), phytogeography<br />

(Billings 1978; Harper et al. 1978) or dominant tree and<br />

shrub distributions (Wells 1983; West et al. 1978). Most<br />

often these studies have applied colonization-extinction<br />

models (sensu MacArthur and Wilson 1967) to mountain<br />

ranges as “sky islands” within an homogenous<br />

“sagebrush sea” rather than actual landscape patterns of<br />

local plant endemism. While such imagery has popular<br />

appeal, it can mask the reality that low elevation ”sea”<br />

contains many distinctive ecological islands supporting<br />

edaphic plant endemics, contributing significantly to the<br />

overall biodiversity of the Great Basin.<br />

This is not simply a matter of scale because ecological<br />

islands in the valleys are more-or-less equivalent<br />

counterparts in both area and isolation to subalpine and<br />

alpine habitats on ridges and peaks. The principal difference<br />

with respect to vulnerability to climate change is<br />

that montane and higher elevation habitats possess a<br />

nearly continuous array of microclimates produced by<br />

variations in slope and aspect across a wide range of<br />

elevation. In contrast, valley habitats have much less<br />

variation in either slope or aspect within a much narrower<br />

elevation range. Thus, lower elevation plant endemics<br />

are likely to have fewer ecological options as<br />

their bioclimatic envelope shifts. This greater vulnerability<br />

to climate change is compounded by the fact that<br />

valley habitats are also more susceptible to invasion by<br />

non-native plants and habitat modification or destruction<br />

by human activities (e.g. transportation or energy<br />

corridors and off-road recreation).<br />

Conceptual Model<br />

Based on the foregoing analysis, I am proposing a<br />

conceptual model for assessing the potential effects of<br />

climate change on rare plants in the Great Basin of Nevada;<br />

this model also may have general application<br />

throughout areas of the west with similar “basin-andrange”<br />

topography (Figure 3a). The model begins with a<br />

generic Great Basin landscape of playas, sand dunes,<br />

and terraces comprised of older lake sediments on valley<br />

fill. At the base of the bounding range, the valley fill<br />

is overlain by alluvial fan deposits. The bounding range<br />

98


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

in this model is comprised of bedded carbonates overlain<br />

by more erosion-resistant quartzite, a typical landscape<br />

throughout much of eastern Nevada. In other parts<br />

of the Great Basin, the ranges may be comprised of volcanic<br />

rocks or, locally, intrusive granitic rocks.<br />

The valley currently provides suitable habitat for<br />

endemic plants that are typically restricted to specialized<br />

edaphic conditions, such as playa edges, the periphery<br />

of sand dunes, or specific microhabitats within older<br />

lake sediments. Montane endemics are restricted to carbonate<br />

or siliceous rock types or, in some cases, are<br />

more restricted to habitats characterized by coarse materials<br />

weathered from various substrates, such as gravels,<br />

angular slates, talus, or scree. At the highest elevations<br />

are the mountaintop endemics in subalpine-alpine habitats<br />

or, in the lower mountain ranges, on ridge tops<br />

within lower vegetation zones (Figure 3b).<br />

As climate changes, populations of those endemic<br />

plants with the most restrictive and least common habitat<br />

constraints are likely to shrink in areal extent and<br />

become more isolated from one another. This is most<br />

likely to happen in the valleys, where many of the rarest<br />

endemics occur. But all plants restricted to highly specialized<br />

habitats, such as mountain tops or carbonate<br />

substrates, are vulnerable when physiological limits are<br />

exceeded as their bioclimatic envelope shifts to unsuitable<br />

habitats. Less specialized endemics may be less<br />

susceptible to habitat shifts but are still likely to decrease<br />

in extent since less area is available at higher elevation<br />

(Figure 3c). The eventual outcome of this scenario<br />

is likely to be extirpation of populations and eventual<br />

species extinctions throughout the landscape, with<br />

the highest rates likely in the valleys where the greatest<br />

edaphic specializations occur (Figure 3d). As noted<br />

3a.<br />

3b.<br />

3c. 3d.<br />

Figure 3. Conceptual model of the effects of climate change on endemic plants in the Great Basin. a) Generalized<br />

landscape typical of the Great Basin in eastern Nevada showing a playa lake, sand dunes, and Pliocene-Pleistocene<br />

lake sediments on the valley floor, overlain by alluvial, and a fault-block mountain range comprised of a band of<br />

carbonate sediments overlain by erosion resistant quartzite that forms the ridge; b) the current landscape occupied<br />

by endemic plant populations restricted to playa edges, sand dunes, carbonate rocks, montane plants that occur on<br />

both carbonates and quartzite, and higher elevation plants; c) as climate change progresses, populations of plants<br />

restricted to specialized habitats contract in place, while those less specialized migrate upward or onto more suitable<br />

aspects as their bioclimatic envelope shifts; d) eventually, populations of plants on highly specialized habitats<br />

are extirpated and the taxa go extinct while other species continue to migrate upward where less habitat area is<br />

available.<br />

99


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

earlier, however, the physiological limits of most species<br />

are unknown and may not be exceeded. Or species<br />

may be able to adapt at a pace commensurate with the<br />

rate of climate change. Alternatively, climate change<br />

may facilitate the migration of competitors into the specialized<br />

habitats of these species. Many uncertainties<br />

remain not only about the parameters of climate change<br />

itself, but also about its potential effects on rare plant<br />

species.<br />

Conservation Strategies<br />

We cannot predict with any certainty which of these<br />

33 plant taxa, or for that matter the roughly 150 additional<br />

G2 taxa (those with fewer than 20 known occurrences),<br />

will be most affected by climate change. Nevertheless,<br />

several conservation actions can be identified to<br />

better ensure the long-term conservation of plant diversity<br />

in the Great Basin. Julius and others (2008) identified<br />

adaptation options with an overall goal of maximizing<br />

resilience to climate change and noted that establishing<br />

current baselines, identifying thresholds, and monitoring<br />

for changes will be essential elements of any adaptation<br />

approach.<br />

Clearly, inventory and monitoring are critical elements<br />

of a conservation strategy for these rare plants<br />

Yet only 23 percent of the 167 reported populations of<br />

the 33 rarest taxa have even been visited in the past decade,<br />

and population estimates are available for fewer<br />

than half of these populations (USFWS, unpublished<br />

data). Quantitative monitoring programs are in place, or<br />

in progress, for only five of the taxa. Without a concerted<br />

effort to establish baselines and a commitment to<br />

long-term monitoring, many populations could vanish<br />

without our knowledge, potentially compromising the<br />

long-term viability of species and certainly resulting in a<br />

loss of genetic resources.<br />

To inventory and monitoring, we can add ex situ approaches<br />

as an, admittedly less than ideal, but necessary<br />

element at least for genetic conservation. Ex situ conservation<br />

has long been regarded as an option of last resort<br />

among plant conservationists, in part, because of concern<br />

that it might be viewed as an acceptable alternative<br />

to the conservation of wild habitats. In the face of climate<br />

change, however, many botanists now recognize<br />

that ex situ conservation has a place among a portfolio<br />

of scientifically based techniques that support the primary<br />

objective of retaining plant diversity in the wild.<br />

Such techniques are requisite for restoration and relocation<br />

actions, especially when integrated with regional<br />

conservation for both ecosystems and suites of species<br />

(Guerrant and Pavlik 1997; Maunder et al. 2004; Pavlik<br />

1996).<br />

Fortunately, both the infrastructure and successful<br />

models for comprehensive and integrated approaches<br />

for ex situ conservation of plant genetic resources exist.<br />

The Center for <strong>Plant</strong> Conservation (CPC; www.center<br />

for plantconservation.org), is dedicated solely to preventing<br />

the extinction of America’s imperiled native<br />

flora. Hosted at the Missouri Botanical Garden in St.<br />

Louis, Missouri, CPC coordinates a network of 33 participating<br />

institutions throughout the country which<br />

maintain plant material (seeds, cuttings, etc.) of the most<br />

imperiled plant species in their region as part of the National<br />

Collection of Endangered <strong>Plant</strong>s, totaling some<br />

700 species.<br />

Representation of the rarest plants of the Great Basin<br />

in the National Collection, however, is poor. Seeds of<br />

only two of the 33 taxa in this study, Eriogonum argophyllum<br />

and Eriogonum ovalifolium var. williamsiae,<br />

are in long-term conservation storage at participating<br />

institutions at the Red Butte Garden in Salt Lake City,<br />

<strong>Utah</strong>, and the Berry Botanical Garden in Portland, Oregon,<br />

respectively. Among the 11 participating institutions<br />

in the western United States, the Berry Botanical<br />

Garden has the most comprehensive collection with 53<br />

taxa conserved, although only a few of these are from<br />

the Great Basin, as delineated here, and it is uncertain<br />

how representative these samples are of the full range of<br />

genetic diversity among these few taxa. The Red Butte<br />

Garden, the designated primary seed repository for the<br />

Great Basin has 22 taxa conserved, most of which are<br />

plants endemic to <strong>Utah</strong>.<br />

Seed collection and long-term storage is also coordinated<br />

by the Bureau of Land Management’s Seeds of<br />

Success project established in 2001 in partnership with<br />

the Royal Botanic Gardens, Kew (http://www.nps.gov/<br />

plants/sos/) to collect, conserve, and develop native<br />

plant materials for stabilizing, rehabilitating and restoring<br />

lands in the United States. This partnership has now<br />

grown to include many partners who have collected<br />

over 6,689 seed accessions. While the focus of Seeds of<br />

Success is on common species, it nevertheless provides<br />

a useful model for a comprehensive, landscape-based<br />

program of targeted seed collection.<br />

CONCLUSIONS<br />

While climate change poses ecosystemic challenges<br />

to many species, narrowly distributed and highly specialized<br />

taxa are particularly at risk. Among the rarest<br />

plants in Great Basin of Nevada, the majority of those<br />

likely to be at greatest risk are restricted to azonal edaphic<br />

habitats in the valleys. A comprehensive and integrated<br />

program of adaptation options is essential to<br />

maximize the resilience of their ecosystems to change.<br />

In particular, inventory, monitoring, and ex situ conservation<br />

are needed to ensure that baseline data are available<br />

against which demographic changes in these taxa<br />

can be evaluated and to ensure that genetic resources<br />

representative of the diversity within these taxa are conserved.<br />

Conservation of the full range of genetic diver-<br />

100


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

sity will enhance the capability of future botanists to<br />

conserve these species in the wild, whether through<br />

population enhancement or translocation to new suitable<br />

sites.<br />

ACKNOWLEDGMENTS<br />

This analysis would not have been possible without<br />

the extended efforts over the past half-century to document<br />

the flora of the Great Basin. The contributors to<br />

this effort are too numerous to list but their names reappear<br />

through the citations herein and in many of the<br />

specific epithets. The Nevada Natural Heritage Program<br />

plays an indispensable role in maintaining data on rare<br />

plant species. Thanks especially to my reviewers, Dr.<br />

Alan de Queiroz, Dr. Bruce Pavlik, and Dr. James<br />

Morefield; their comments helped clarify my thinking<br />

and made a substantial improvement in the manuscript.<br />

Finally, thanks to the Nevada Fish and Wildlife Office<br />

of the U.S. Fish and Wildlife Office for allowing me<br />

time to complete this work. The findings and conclusions<br />

in this article are those of the author and do not<br />

necessarily represent the views of the U.S. Fish and<br />

Wildlife Service.<br />

REFERENCES<br />

Aitken, M., D.W. Roberts, and L.M. Shultz. 2007.<br />

Modeling distributions of rare plants in the Great Basin,<br />

western North America. Western North American Naturalist.<br />

67(1):26–38.<br />

Al-Shehbaz, I.A. and M.D. Windham. 2007. New or<br />

noteworthy North American Draba (Brassicaceae). Harvard<br />

Papers in Botany. 12(2):409–419.<br />

Anderson, S., M. White, and D. Atwood. 1991.<br />

Humboldt National Forest sensitive plant field guide.<br />

Ogden, UT: Department of Agriculture, U.S. Forest Service:<br />

unpaginated.<br />

Barneby, R.C. 1989. Trifolium andinum var. podocephalum.<br />

In: Cronquist, A.; Holmgren, A.H.; Holmgren,<br />

N.H.; Reveal, J.L.; Holmgren, P.K. Intermountain<br />

Flora: vascular plants of the Intermountain West, U.S.A.<br />

Volume Three, Part II (Fabales). Bronx, NY: The New<br />

York Botanical Garden: 217–218.<br />

Beatley, J. 1969. Biomass of desert winter annual<br />

plant populations in southern Nevada. Oikos. 20(2):<br />

261–273.<br />

Beatley, J. 1975. Climates and vegetation pattern<br />

across the Mojave/Great Basin Desert transition of<br />

southern Nevada. American Midland Naturalist. 93:<br />

53–70.<br />

Beever, E., P.F. Brussard, and J. Berger. 2003. Patterns<br />

of apparent extirpation among isolated populations<br />

of pikas (Ochotona princeps) in the Great Basin. Journal<br />

of Mammalogy. 84:37–54.<br />

Billings, W.D. 1949. The shadscale vegetation zone<br />

of Nevada and eastern California in relation to climate<br />

and soils. American Midland Naturalist. 42(1):87–109.<br />

Billings, W.D. 1950. Vegetation and plant growth as<br />

affected by chemically altered rocks in the western<br />

Great Basin. Ecology. 31:62–74.<br />

Billings, W.D. 1978. Alpine phytogeography across<br />

the Great Basin. In: Harper, K.T.; Reveal, J.L. Intermountain<br />

biogeography: a symposium. Great Basin<br />

Naturalist Memoirs 2, Provo, UT: Brigham Young University<br />

Printing Service: 105–118.<br />

Bonfils, C., B.D. Santer, D,W, Pierce, H.G. Hidaloo,<br />

B. Govindasamy, T. Das, T.P. Barnett, D.R. Cayan, C.<br />

Doutriaux, A.W. Wood, A. Mirin, and T. Nozawa.<br />

2008. Detection and attribution of temperature changes<br />

in the mountainous Western United States. Journal of<br />

Climate. 21:6404–6424.<br />

Brown, J.H. 1971. Mammals on mountaintops: nonequilibrium<br />

insular biogeography. American Naturalist.<br />

105:467–478.<br />

Brown, J.H. 1978. The theory of insular biogeography<br />

and the distribution of boreal birds and mammals.<br />

In: Harper, K.T.; Reveal, J.L. Intermountain biogeography:<br />

a symposium. Great Basin Naturalist Memoirs 2,<br />

Provo, UT: Brigham Young University Printing Service:<br />

209–227.<br />

Cayan, D.R., S.A. Kammerdiener, M.D. Dettinger, J.<br />

M. Caprio, and D.H. Peterson. 2001. Changes in the<br />

onset of spring in the western United States. Bulletin of<br />

the American Meteorological <strong>Society</strong>. 82:399–415.<br />

Charlet, D. 1996. Atlas of Nevada conifers: a phytogeographic<br />

reference. Reno, NV. University of Nevada<br />

Press: 320 p.<br />

Committee on Environment and Natural Resources.<br />

2008. Scientific assessment of global change on the<br />

United States. Washington, D.C., U.S. Climate Change<br />

Science Program, National Science and Technology<br />

Council: 261 p.<br />

Comstock, J.P. and J.R. Ehleringer. 1992. <strong>Plant</strong> adaptation<br />

in the Great Basin and Colorado Plateau. Great<br />

Basin Naturalist. 52(3):195–215.<br />

Crowder, D.F. and M.F. Sheridan. 1972. Geologic<br />

map of the White Mountain Peak Quadrangle, Mono<br />

County, California. U.S. Geologic Survey Report GQ–<br />

1012.<br />

Cubashi, U, G.A. Meehl, and G.J. Boer. 2001. Projections<br />

of future climate change. In: Houghten, J.T.,<br />

Ding, Y.; Griggs, D.J.; Noguer, M.; van der Linden,<br />

P.J.; Da, X.; Maskell, K.; Johnson, C.A., eds. Climate<br />

change 2001: the scientific basis. Contributions of<br />

Working Group I to the Third Assessment Report of the<br />

Intergovernmental Panel on Climate Change. Cambridge,<br />

Cambridge University Press: [Online] Available:<br />

http://www.grida.no/publications/other/ipcc%5Ftar/?<br />

src=/CLIMATE/IPCC_TAR/wg1/109.htm. [March 13,<br />

2009].<br />

101


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Danley, D.M. 1985. A new Penstemon (Scrophulariaceae)<br />

from northwestern Nevada. Brittonia. 37:<br />

321–324.<br />

Fautin, R.W. 1946. Biotic communities of the northern<br />

desert shrub biome in western <strong>Utah</strong>. Ecological<br />

Monographs. 16(4):251–310.<br />

Fischlin, A., G.F. Midgley, J.T. Price, R. Leemans,<br />

B. Bopal, C. Turley, M.D.A. Rounsevell, O.P. Dube, J.<br />

Tarazona, and A.A. Velichko. 2007. Ecosystems, their<br />

properties, goods, and services. In: Parry, M.L.; Canziani,<br />

O.F., Palutikof, J.P.; van der Linden, P.J.; and<br />

Hanson, C.E., eds. Climate change 2007: Impacts, adaptation<br />

and vulnerability. Contributions of Working<br />

Group II to the Fourth Assessment Report of the Intergovernmental<br />

Panel of Climate Change. Cambridge,<br />

Cambridge University Press: [Online] Available:<br />

http://www.ipcc.ch/ipccreports/ar4-wg2.htm. [May 26,<br />

2009]<br />

Forbis, T. 2007. White River rare plant conservation<br />

strategy. Unpublished report prepared by The Nature<br />

Conservancy, Reno, Nevada. On file at Nevada Fish and<br />

Wildlife Office, U.S. Fish and Wildlife Service, Reno,<br />

NV. 20 p.<br />

Goodrich, S. 1979. Status report on Eriogonum<br />

holmgrenii. Unpublished report prepared for the U.S.<br />

Fish and Wildlife Service, Portland, OR. On file at Nevada<br />

Fish and Wildlife Office, U.S. Fish and Wildlife<br />

Service, Reno, NV. 5 p.<br />

Grabherr, G., M. Gottfried, A. Gruber, and H. Pauli.<br />

1995. Patterns and current changes in alpine plant diversity.<br />

In: Chapin, F.S.; Körner, C., eds. Arctic and Alpine<br />

Biodiversity. Berlin, Heidelberg: Springer Verlag. p.<br />

167–181.<br />

Grayson, D.K. 2000. Mammalian responses to Middle<br />

Holocene climatic change in the Great Basin of the<br />

western United States. Journal of Biogeography.<br />

27:181–192.<br />

Grimes, J.W. 1984. Notes on the flora of Leslie<br />

Gulch, Malheur County, Oregon. Madroño 31:69–85.<br />

Guerrant, E.O., Jr. and B.M. Pavlik. 1997. Reintroduction<br />

of rare plants: genetics, demography, and the<br />

role of ex situ methods. In: Fiedler, P.L.; Jain, S.K.,<br />

eds. Conservation Biology: the theory and practice of<br />

nature conservation, preservation, and management.<br />

Second Edition. London: Chapman and Hall:80–108.<br />

Halloy, S.R.P. and A.F. Mark. 2003. Climate-change<br />

effects on alpine plant biodiversity: a New Zealand perspective<br />

on quantifying threat. Arctic, Antarctic, and<br />

Alpine Research. 35(2):248–254.<br />

Harper, K.T. and J.L. Reveal. 1978. Intermountain<br />

biogeography: a symposium. Great Basin Naturalist<br />

Memoirs 2, Provo, UT: Brigham Young University<br />

Printing Service: 268 pp.<br />

Heil, K.D. and J.M. Porter. 1994. Sclerocactus<br />

(Cactaceae): a revision. Haseltonia. 2:20–46.<br />

Hidy, G.M. and H.E. Klieforth. 1990. Atmospheric<br />

processes affecting the climate of the Great Basin. In:<br />

Osmond, C.B.; Pitelka, L.F.; Hidy, G.M., eds. <strong>Plant</strong><br />

biology of the Basin and Range. Ecological Studies 80,<br />

Springer-Verlag, Berlin. 19–45.<br />

Holland, R.F. 1995. Current knowledge and conservation<br />

status of Primula capillaris Holmgren & Holmgren<br />

(Primulaceae), Ruby Mountains primrose, in Nevada.<br />

Unpublished report prepared for Nevada Natural<br />

Heritage Program, Carson City, Nevada, and Nevada<br />

Fish and Wildlife Office, U.S. Fish and Wildlife Service,<br />

Reno, Nevada. On file at Nevada Fish and Wildlife<br />

Office, U.S. Fish and Wildlife Service, Reno, NV. 19 p.,<br />

plus appendices.<br />

Holmgren, A.H. 1954. A new Lewisia from Nevada.<br />

Leaflets of Western Botany. 7:135–137.<br />

Holmgren, N.H. 1972a. <strong>Plant</strong> geography of the Intermountain<br />

Region. In: Cronquist, A.; Holmgren, A.H.;<br />

Holmgren, N.H.; Reveal, J.L. Intermountain Flora: vascular<br />

plants of the Intermountain West, U.S.A. Volume<br />

One. Bronx, NY: The New York Botanical Garden: 77–<br />

161.<br />

Holmgren, N.H. 1972b. Five new species of Castilleja<br />

(Scrophulariaceae) from the Intermountain Region.<br />

Bulletin of the Torrey Botanical Club. 100(2):83–<br />

93.<br />

Holmgren, N.H. 1979. New penstemons (Scrophulariaceae)<br />

from the intermountain region. Brittonia. 31:<br />

217–242.<br />

Holmgren, N.H. 1987. Two new species of Potentilla<br />

(Rosaceae) from the Intermountain Region of western<br />

U.S.A. Brittonia. 39(3):340–344.<br />

Holmgren, N.H. 1992. Two new species of Viola<br />

(Violaceae) from the Intermountain West, U.S.A. Brittonia.<br />

44(3):300–305.<br />

Holmgren, N.H. 1998. Two new species of Penstemon<br />

(Scrophulariaceae: Sect. Saccanthera) from Nevada,<br />

U.S.A. Brittonia. 50(2):159–164.<br />

Holmgren, N.H. and A.H. Holmgren. 1974. Three<br />

new species from the Great Basin. Brittonia. 26(3):309–<br />

315.<br />

Holmgren, N.H. and P.K. Holmgren. 2002. New<br />

Mentzelias (Loasaceae) from the Intermountain Region<br />

of the western United States. Systematic Botany. 27<br />

(4):747-762.<br />

Hunter, K.B. and R.E. Johnson. 1983. Alpine flora of<br />

the Sweetwater Mountains, Mono County, California.<br />

Madroño 30(4):89–105.<br />

Jenny. 1941. Factors of soil formation. New York:<br />

McGraw-Hill. 109 pp<br />

Joyal, E. 1986. A new species of Collomia<br />

(Polemoniaceae) from the Great Basin. Brittonia. 38<br />

(3):243–248.<br />

Julius; S.H., J.M. West, G.M. Blate, J.S. Baron, B.<br />

Griffith, L.A. Joyce, P. Kareiva, B.D. Keller, M.A.<br />

102


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Palmer, C.H. Peterson, and J.M. Scott. 2008. Executive<br />

Summary. In: Julius, S.H.; West, J.M. eds. Preliminary<br />

review of adaptation options for climate-sensitive ecosystems<br />

and resources. A report by the U.S. Climate<br />

Change Science Program and the Subcommittee on<br />

Global Change Research. Washington, D.C., U.S. Environmental<br />

Protection Agency: p. 1–1 to 1–6.<br />

Kartesz, J.T. 1987. A flora of Nevada. Reno, NV:<br />

University of Nevada, Reno. 1,729 p. Dissertation.<br />

Kleinhampl, F.J. and J.L. Ziony. 1985. Geology of<br />

northern Nye County, Nevada. Bulletin 93a. Reno, NV:<br />

University of Nevada, Reno, Nevada Bureau of Mines,<br />

Mackay School of Mines. 172 p.<br />

Knight, T. 1988. Status report for Penstemon floribundus<br />

Danley. Unpublished report prepared by the Nevada<br />

Natural Heritage Program for the Winnemucca<br />

District, Bureau of Land Management, Winnemucca,<br />

NV. On file at Nevada Fish and Wildlife Office, U.S.<br />

Fish and Wildlife Service, Reno, NV. 23 p., plus appendices.<br />

Kruckeberg, A.R. 1986. An essay: The stimulus of<br />

unusual geologies for plant speciation. Systematic Botany.<br />

11:455–463.<br />

Kruckeberg, A.R. and D. Rabinowitz. 1985. Biological<br />

aspects of endemism in higher plants. Annual Review<br />

of Ecology and Systematics. 16:447–479.<br />

Lawlor, T.E. 1998. Biogeography of Great Basin<br />

mammals: paradigm lost? Journal of Mammalogy<br />

79:1111–1130.<br />

Lenoir, J., J.C. Gégout, P.A. Marquet, P. de Ruffray,<br />

and H. Brisse. 2008. A significant upward shift in plants<br />

species optimum elevation during the 20 th century. Science.<br />

320:1768–1771.<br />

Lesica, P. and B. McCune. 2004. Decline of arcticalpine<br />

plants at the southern margin of their range following<br />

a decade of climatic warming. Journal of Vegetation<br />

Science. 15:679–690.<br />

MacArthur, R.H. and E.O. Wilson. 1967. The theory<br />

of island biogeography. Princeton, NJ. Princeton University<br />

Press: 203 p.<br />

Major, J. 1951. A functional, factorial approach to<br />

plant ecology. Ecology. 32(3):392–412.<br />

Marchand, D.E. 1973. Edaphic control of plant distribution<br />

in the White Mountains, eastern California.<br />

Ecology, 54:233–250.<br />

Mason, H.L. 1946a. The edaphic factor in narrow<br />

endemism. I. The nature of environmental differences.<br />

Madroño 8:209–226.<br />

Mason, H.L. 1946b. The edaphic factor in narrow<br />

endemism. II. The geographic occurrence of plants of<br />

highly restricted patterns of distribution. Madroño 8:<br />

241–272.<br />

Maunder, M., E.O. Guerrant, Jr., K. Havens, and<br />

K.W. Dixon. 2004. Realizing the full potential of Ex<br />

Situ contributions to global plant conservation. In:<br />

Guerrant, E.O., Jr.; Havens, K., Maunder, M., eds. Exsitu<br />

<strong>Plant</strong> Conservation. Washington: Island Press:<br />

389–418.<br />

Meyer, S. 1978. Some factors governing plant distributions<br />

in the Mojave-Intermountain transition zone. In:<br />

Harper, K.T.; Reveal, J.L. 1978. Intermountain biogeography:<br />

a symposium. Great Basin Naturalist Memoirs 2,<br />

Provo, UT: Brigham Young University Printing Service:<br />

197–207.<br />

Mooney, H.A. 1966. Influence of soil type on the<br />

distribution of two closely related species of Erigeron.<br />

Ecology, 47:950–958.<br />

Mooney, H.A., G. St. Andre, and R.D. Wright. 1962.<br />

Alpine and subalpine vegetation patterns in the White<br />

Mountains of California. American Midland Naturalist.<br />

68:257–273.<br />

Morefield, J.D. 1992. Spatial and ecologic segregation<br />

of phytogeographic elements in the White Mountains<br />

of California and Nevada. Journal of Biogeography.<br />

19(1):33–50.<br />

Morefield, J.D. 1995. Current knowledge and conservation<br />

status of Eriogonum tiehmii Reveal (Polygonaceae),<br />

Tiehm buckwheat. Unpublished report prepared<br />

by the Nevada Natural Heritage Program, Carson City,<br />

Nevada. On file at Nevada Fish and Wildlife Office,<br />

U.S. Fish and Wildlife Service, Reno, NV. 21 p., plus<br />

appendices.<br />

Morefield, J.D. 1997. Current knowledge and conservation<br />

status of Arabis falcifructa Rollins (Brassicaceae),<br />

the Elko rockcress. Unpublished report prepared<br />

by the Nevada Natural Heritage Program, Carson City,<br />

Nevada, for the Bureau of Land Management, Reno,<br />

Nevada. On file at Nevada Fish and Wildlife Office,<br />

U.S. Fish and Wildlife Service, Reno, NV. 28 p., plus<br />

appendices.<br />

Morefield, J.D., ed. 2001. Nevada rare plant atlas.<br />

Unpublished report compiled by the Nevada Natural<br />

Heritage Program, Carson City, Nevada, for the U.S.<br />

Fish and Wildlife Service, Portland, OR, and Reno, NV.<br />

[Online]. Available: http://heritage.nv.gov/atlas/<br />

atlas.html). [March 12, 2009].<br />

Morefield, J.D. 2003. Current knowledge and conservation<br />

status of Arabis ophira Rollins (Brassicaceae),<br />

the Ophir rockcress. Unpublished status report prepared<br />

by the Nevada Natural Heritage Program, Carson City,<br />

NV. On file at Nevada Fish and Wildlife Office, U.S.<br />

Fish and Wildlife Service, Reno, NV. 25 p., plus appendices.<br />

Mote, P.W., A.F. Hamlet, M.P. Clark., and D.P. Lettermaier.<br />

2005. Declining mountain snowpack in western<br />

North America. Bulletin of the American Meteorological<br />

<strong>Society</strong>. 86:39–49.<br />

Mozingo, H.M and M. Williams. 1980. Threatened<br />

and endangered plants of Nevada: an illustrated manual.<br />

Portland, OR, U.S. Fish and Wildlife Service, and U.S.<br />

103


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Bureau of Land Management, Reno, NV: 268 p.<br />

Murphy, D.D. and S.B. Weiss. 1992. Effects of climate<br />

change on biological diversity in western North<br />

America: species losses and mechanisms. In: Peters,<br />

R.L.; Lovejoy, T.E., eds. Global warming and biological<br />

diversity. Castleton, NY: Hamilton Printing: Chapter 26.<br />

Nachlinger, J., K. Sochi, P. Comer, G. Kittel, and D.<br />

Dorfman. 2001. Great Basin: an ecoregion-based conservation<br />

blueprint. Reno, Nevada: The Nature Conservancy.<br />

159 p., plus appendices.<br />

NatureServe. 2009. NatureServe Explorer. [Online]<br />

Available: http://www.natureserve.org/explorer/<br />

ranking.htm. [May 26, 2009]<br />

Parmesan, C. 2006. Ecological and evolutionary responses<br />

to recent climate change. Annual Review of<br />

Ecology and Systematics. 37:637–669.<br />

Pavlik, B.M. 1989. Phytogeography of sand dunes<br />

in the Great Basin and Mojave Deserts. Journal of Biogeography.<br />

16(3):227–238.<br />

Pavlik, B.M. 1996. A framework for defining and<br />

measuring success during reintroduction of endangered<br />

plants. In: Falk, D; Millar, C.; Olwell, P., eds. Restoring<br />

diversity: strategies for reintroduction of endangered<br />

plants. Washington: Island Press:127–156.<br />

Peters, R.L. and J.D.S. Darling. 1985. The greenhouse<br />

effect and nature reserves. BioScience. 35<br />

(11):707–717.<br />

Pritchett, D.W. and R. Patterson. 1998. Morphological<br />

variation in California alpine Polemonium species.<br />

Madroño 45(3):200–209.<br />

Rajakaruna, N. 2004. The edaphic factor in the origin<br />

of plant species. International Geology Review. 46:471–<br />

478.<br />

Ramsey, D. and L. Shultz. 2004. Evaluating the geographic<br />

distribution of plants in <strong>Utah</strong> from the Atlas of<br />

Vascular <strong>Plant</strong>s in <strong>Utah</strong>. Western North American<br />

Naturalist. 64(4):421–432.<br />

Reveal, J.L. 1965. A new alpine Eriogonum from<br />

Nevada. Leaflets of Western Botany. 10:183–186.<br />

Reveal, J.L. 1972. New species and combinations in<br />

Eriogonum. Phytologia. 23(1):168–172.<br />

Reveal, J.L. 1979. Distribution and phylogeny of<br />

Eriogonoideae (Polygonaceae). In: Harper, K.T.; Reveal,<br />

J.L. 1978. Intermountain biogeography: a symposium.<br />

Great Basin Naturalist Memoirs 2, Provo, UT:<br />

Brigham Young University Printing Service: 169–190.<br />

Reveal, J.L. 1981. Notes on endangered buckwheats<br />

(Eriogonum: Polygonaceae) with three newly described<br />

from the United States. Brittonia. 33(3):441–448.<br />

Reveal, J.L. 1985. New Nevada entities and combinations<br />

in Eriogonum (Polygonaceae). Great Basin<br />

Naturalist. 45(2):276–280.<br />

Reveal, J.L. 2004a. Johanneshowellia (Polygonaceae:<br />

Eriogonoideae), a new genus from the Intermountain<br />

West. Brittonia. 56(4):299–306.<br />

Reveal, J.L. 2004b. New entities in Eriogonum<br />

(Polygonaceae: Eriogonoideae). Phytologia. 86(3):121–<br />

159.<br />

Reveal, J.L. and J. Beatley. 1971. A new Penstemon<br />

(Scrophulariaceae) and Grindelia (Asteraceae) from<br />

southern Nye County, Nevada. Bulletin of the Torrey<br />

Botanical Club. 98:332–335.<br />

Reveal, J.L., J. Reynolds, and J. Picciani. 2002.<br />

Eriogonum diatomaceum (Polygonaceae: Eriogonoideae),<br />

a New Species from Western Nevada, U.S.A.<br />

Novon. 12:87–89.<br />

Rundel, P.W., A.C. Gibson, and M.R. Sharifi. 2008.<br />

The alpine flora of the White Mountains, California.<br />

Madroño 55(3):202–215.<br />

Shreve, F. 1942. The desert vegetation of North<br />

America. The Botanical Review. 8(4):195–247.<br />

Smith, F. 1994. Status report for Frasera gypsicola<br />

(Barneby 1942) D. Post. Unpublished report prepared<br />

for Nevada Natural Heritage Program, Carson City, Nevada,<br />

and Nevada Fish and Wildlife Office, U.S. Fish<br />

and Wildlife Service, Reno, Nevada. On file at Nevada<br />

Fish and Wildlife Office, U.S. Fish and Wildlife Service,<br />

Reno, NV. 12 p., plus appendices.<br />

Snyder, M.A. and L.C. Sloan. 2005. Transient future<br />

climate over the western United States using a regional<br />

climate model. Earth Interactions 9, Paper No. 11:1-21.<br />

[Online]. Available: http://EarthInteractions.org. [March<br />

12, 2009].<br />

Spahr, R., L. Armstrong, D. Atwood, and M. Rath.<br />

1991. Atwood. Threatened, Endangered, and Sensitive<br />

Species of the Intermountain Region. Ogden, UT: Department<br />

of Agriculture, U.S. Forest Service: unpaginated.<br />

Tiehm, A. and B. Ertter. 1984. Potentilla basaltica<br />

(Rosaceae), a new species from Nevada. Brittonia. 36<br />

(3):228–231.<br />

Tiehm, A. and P.K. Holmgren. 1991. A new variety<br />

of Draba oreibata (Brassicaceae) from Nevada, U.S.A.<br />

Brittonia. 43(1):20–23.<br />

Tiehm, A. and L. Shultz. 1985. A new Haplopappus<br />

(Asteraceae: Astereae) from Nevada. Brittonia. 37<br />

(2):165–168.<br />

Tingley, J.V. and K.A. Pizarro. 2000. Traveling<br />

America’s loneliest road: a geologic and natural history<br />

tour through Nevada along U.S. Highway 50. Nevada<br />

Bureau of Mines and Geology, Special Publication<br />

26, Mackay School of Mines, University of Nevada,<br />

Reno.<br />

Tschanz, C.M. and E.H. Pampeyan. 1970. Geology<br />

and mineral resources of Lincoln County, Nevada. Bulletin<br />

73. Reno, NV: University of Nevada, Reno, Nevada<br />

Bureau of Mines, Mackay School of Mines. 187 p.<br />

Van de Ven, C.M., S.B. Weiss, and W.G. Ernst.<br />

2007. <strong>Plant</strong> species distributions under present conditions<br />

and forecasted for warmer climates in an arid<br />

104


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

mountain range. Earth Interactions. 11:1-33. [Online].<br />

Available: http://EarthInteractions.org. [March 12,<br />

2009].<br />

Wagner, F.H., ed. 2003. Preparing for a changing<br />

climate – the potential consequences of climate variability<br />

and change. Rocky Mountain/Great Basin regional<br />

climate-change assessment. A report of the Rocky<br />

Mountain/Great Basin Regional Assessment Team for<br />

the U.S. Global Change Research Program. Logan, UT.<br />

<strong>Utah</strong> State University. 240 p.<br />

Walther, G., S. Beißner, and C.A. Burga. 2005.<br />

Trends in the upward shift of alpine plants. Journal of<br />

Vegetation Science. 16:541–548.<br />

Weixelman, D. and N.D. Atwood. Undated. Toiyabe<br />

National Forest sensitive plant field guide. Ogden, UT.<br />

U.S. Department of Agriculture, U.S. Forest Service:<br />

123 p.<br />

Wells, P.V. 1983. Paleobiogeography of montane<br />

islands in the Great Basin since the last glaciopluvial.<br />

Ecological Monographs. 53(4):341–382.<br />

Welsh, S.L. and E. Neese. 1980. A new species of<br />

Cymopterus (Umbelliferae) from the Toiyabe Range,<br />

Lander County, Nevada. Madroño. 27(2):97–100.<br />

Welsh, S.L. and K.H. Thorne. 1985. New Sclerocactus<br />

(Cactaceae) from Nevada. Great Basin Naturalist. 45<br />

(3):553–555.<br />

West, N.E., R.J. Tausch, K.H. Rea, and P.T. Tueller.<br />

1978. In: Harper, K.T.; Reveal, J.L. Intermountain biogeography:<br />

a symposium. Great Basin Naturalist Memoirs<br />

2, Provo, UT: Brigham Young University Printing<br />

Service: 119–136<br />

Wharton, R.A., P.E. Wigand, M.R. Rose, R.L.<br />

Reinhardt, D.A. Mouat, H.E. Kleiforth, N.L. Ingraham,<br />

J.O. Davis, C.A. Fox, and J.T. Ball. 1990. The North<br />

American Great Basin: a sensitive indicator of climatic<br />

change. In: Osmond C.B.; Pitelka, L.F.; Hidy, G.M.,<br />

eds. <strong>Plant</strong> biology of the Basin and Range. Ecological<br />

Studies 80. New York, NY: Springer-Verlag: 323–359.<br />

Williams, M.J. 1981. Status report on Lewisia<br />

maguirei. Unpublished report prepared for the U.S. Fish<br />

and Wildlife Service, Portland, OR. On file at the Nevada<br />

Fish and Wildlife Office, U.S. Fish and Wildlife<br />

Service, Reno, NV. 19 p.<br />

Wright, R.D. and H.A. Mooney. 1965. Substrateoriented<br />

distribution of Bristlecone pine in the White<br />

Mountains of California. American Midland Naturalist.<br />

73(2):257–284.<br />

105


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Sentry Milkvetch (Astragalus cremnophylax<br />

var. cremnophylax) Update<br />

Janice Busco<br />

Grand Canyon National Park, Grand Canyon, AZ<br />

Abstract. The Grand Canyon endemic Astragalus cremnophylax var. cremnophylax (Sentry milkvetch) was listed as<br />

an endangered species in 1990. There are 725 plants within three populations on the South Rim, all found in shallow<br />

soils upon large flat Kaibab limestone platforms. Habitat specificity and reduced number of plants make Sentry milkvetch<br />

vulnerable to extinction. Recovery plan actions completed in 2008 include seed collection and parking lot removal<br />

to allow habitat restoration and population expansion. Planned recovery plan actions include establishment of<br />

an ex situ population, seed production, and development of techniques for population augmentation and creation of<br />

artificial populations.<br />

The Grand Canyon National Park endemic Astragalus<br />

cremnophylax Barneby var. cremnophylax (Sentry<br />

milkvetch) was listed as an endangered species by the<br />

US Fish and Wildlife Service and protected from trampling<br />

by South Rim sightseers in 1990. There are approximately<br />

725 individuals known in three locations,<br />

all on the South Rim of Grand Canyon. Sentry milkvetch<br />

occurs in shallow, well-drained soils or porous<br />

limestone pavement in crevices and depressions in large<br />

flat Kaibab limestone platforms in unshaded openings in<br />

the pinyon-juniper woodland along the canyon edge.<br />

The underlying bedrock limestone stores water and is<br />

critical to its growth and development (USFWS 2006).<br />

Sentry milkvetch is a small, mat-forming perennial<br />

plant (Figure 1) and has a thick taproot and woody caudex.<br />

Pale purple flowers appear from late April to early<br />

May, with seed set in late May to June. Its tiny seeds<br />

tend to fall in the mat of the plant; therefore the plant<br />

does not spread and remains isolated.<br />

Threats to Sentry milkvetch include small population<br />

size, vulnerability to drought and stochastic events, digging<br />

by ground squirrels and bighorn sheep (Figure 2),<br />

low reproductive capacity, limited seed dispersal, limited<br />

habitat, and reduced genetic diversity and vigor<br />

(Allphin et al. 2005).<br />

Figure 1. A mature Sentry milkvetch with a quarter for<br />

scale.<br />

RECOVERY CRITERIA AND OBJECTIVES<br />

In order to downlist the species, the Sentry Milkvetch<br />

Recovery Plan (USFWS 2006) requires achievement,<br />

maintenance and long-term protection of at least<br />

four viable Sentry milkvetch populations of at least<br />

1,000 individuals each, for a total of at least 4,000 individuals<br />

in the wild. Recovery will be attained when<br />

there are eight viable Sentry milkvetch populations of<br />

1,000 individuals each, with long-term protection.<br />

Figure 2. Bighorn sheep damage to Sentry milkvetch.<br />

106


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

RECOVERY PLAN ACTIONS<br />

Completed Actions 2006-2008<br />

In 2006, recovery plan actions for the species began.<br />

Grand Canyon NP has completed annual monitoring of<br />

the Maricopa Point population each year (2006-2008)<br />

and performed a complete census of the Maricopa Point<br />

population in 2008. We installed permanent photopoints<br />

in all three populations in summer 2008. Grand<br />

Canyon NP and the Arboretum at Flagstaff collected<br />

2600 seeds from 74 individuals in the summer of 2008<br />

(Figure 3). The Arboretum at Flagstaff completed seed<br />

germination studies, initiated greenhouse seed production.<br />

established an ex situ population, and conducted<br />

mycorrhizal studies of the species. Grand Canyon NP<br />

completed parking lot removal, trail rerouting and shuttle<br />

bus stop relocation from Sentry milk-vetch habitat<br />

adjacent to the Maricopa Point population to allow habitat<br />

restoration and population augmentation and expansion<br />

(Figure 4). In addition, two suitable areas for artificial<br />

population establishment were selected in Spring<br />

2008.<br />

Planned Actions 2009-2013<br />

Recovery plan actions planned for 2009-2013 include<br />

restoration of disturbed habitat and completion of<br />

seeding and planting trials at Maricopa Point. Acquisition<br />

of a passive solar greenhouse in 2009 will provide a<br />

dedicated facility for seeding trials, seed and plant production<br />

for introduction trials at Maricopa Point, and<br />

will create an ex situ Sentry milkvetch population.<br />

Through these trials we plan to develop techniques for<br />

creation of artificial populations, increase the number of<br />

individuals at Maricopa Point, and establish new pilot<br />

populations in suitable areas near Maricopa Point<br />

(Figure 5). A discovery survey of the westernmost portions<br />

of the south rim will be completed in 2009. Soil<br />

seed bank studies and ecological observations to determine<br />

pollinators will be conducted in 2009-2010. Additionally,<br />

interpretative materials will be developed and<br />

displayed at the Grand Canyon Visitor Center. Cooperating<br />

agencies in planned recovery actions include<br />

Grand Canyon National Park, US Fish and Wildlife Service,<br />

The Arboretum at Flagstaff, Grand Canyon Association,<br />

National Park Service, Center for <strong>Plant</strong> Conservation,<br />

Northern Arizona University Environmental<br />

Monitoring and Assessment (EMA) and Coconino National<br />

Forest.<br />

LITERATURE CITED<br />

Allphin, L., N. Brian, and T. Matheson. 2005. Reproductive<br />

success and genetic divergence among varieties<br />

of the rare and endangered Astragalus cremnophylax<br />

(Fabaceae) from Arizona, USA. Conservation Genetics<br />

6: 803-821.<br />

Figure 3. Seed collection, summer 2008.<br />

Figure 4. Parking lot removal adjacent to population,<br />

September 2008, readied the site for restoration to be<br />

completed in 2010-<strong>2012</strong>.<br />

Figure 5. Suitable area selected for Sentry milkvetch<br />

artificial population.<br />

107


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 7. A Mason bee (Osmia ribifloris ribifloris) pollinating<br />

Sentry milkvetch.<br />

Figure 6. A Painted lady butterfly visits Sentry milkvetch.<br />

USFWS 2006. Sentry Milk-vetch (Astragalus<br />

cremnophylax Barneby var. cremnophylax Barneby)<br />

Recovery Plan. U.S. Fish and Wildlife Service, Albuquerque,<br />

New Mexico, i-vii + 44pp.<br />

ADDENDUM<br />

This paper reported on the earliest efforts of Grand<br />

Canyon National Park to begin implementation of the<br />

2006 USFWS recovery plan for Sentry milkvetch.<br />

Since my presentation at the University of <strong>Utah</strong> in<br />

2009, four years of intensive work to recover this species<br />

has been completed. Some of this work is documented<br />

in Falk and others (2011) and Busco and others<br />

(2011).<br />

Preliminary pollination studies in spring 2010 established<br />

the identity of three pollinators - two mason<br />

bee (Osmia ribifloris and O. ribifloris ribifloris, Figure<br />

7) and a hoverfly (Syrphidae) (Busco et al. 2011).<br />

We have confirmed the presence of these generalist<br />

pollinators throughout Sentry milkvetch populations<br />

in repeat studies in 2011 and <strong>2012</strong>.<br />

At the time of my presentation in 2009, only 725<br />

individuals of Sentry milkvetch were known. Today<br />

there are an estimated 3552 known naturally-occurring<br />

individuals in wild populations, and 425 plants in<br />

reintroduction areas. The increase in numbers within<br />

wild populations is largely the result of the discovery<br />

of new groups of Sentry milkvetch plants on limestone<br />

fingers above the rim and on lower limestone<br />

levels below the rim during revisits to these populations<br />

in 2010-<strong>2012</strong> (Figure 8), as well as the result of<br />

continued protection of the Maricopa Point population.<br />

While two of the three populations are apparently<br />

stable or increasing in number, the third small<br />

population is increasingly threatened. The area below<br />

the rim of this population has crumbled away and<br />

fallen into the canyon; the few remaining individuals on<br />

a solitary boulder above may likely follow.<br />

<strong>Plant</strong> reintroductions at Maricopa Point began in July<br />

2010 and continue to this date (Figure 9). The first<br />

small planting trial was completed in July 2010 – 5 Sentry<br />

milkvetch plants that were planted from greenhousegrown<br />

plants at that time are all alive today. Seeds were<br />

less successful in that reintroduction - 10 groups of three<br />

seeds each were sown in April 2011 and today one seedling<br />

from this cohort is alive and has reached reproductive<br />

maturity. Of eighty greenhouse-grown plants and<br />

Figure 8. Newly discovered Sentry milkvetch population<br />

in Grand Canyon National Park.<br />

108


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 9. First large-scale reintroduction site for Sentry<br />

milkvetch.<br />

240 seeds sown in the restored parking lot area adjacent<br />

to the Maricopa Point population in July 2011, 51% survive.<br />

Fourteen have set seed on site and produced 58<br />

seedlings. In addition, another 44 milkvetch seedlings<br />

have become established from seed that either washed<br />

in from the adjacent Sentry milkvetch population or<br />

were present in the soil seed bank and germinated. Of<br />

seeds planted in 2011, 15.8% produced seedling plants<br />

that are now alive. In total, there are now 181 sentry<br />

milk-vetch plants growing in this reintroduction site<br />

where habitat beneath a parking lot removed in 2008<br />

was uncovered and restored.<br />

We seeded a second reintroduction area at Maricopa<br />

Point in <strong>2012</strong>. Two additional seeding trials were carried<br />

out at the Maricopa Point reintroduction areas in<br />

July <strong>2012</strong> with 518 seeds. We also tested seeding techniques<br />

and watering regimens for this species that will<br />

provide information for continued large-scare reintroduction<br />

efforts.<br />

Finally, we have received a three-year National Park<br />

Service grant to complete installation and establishment<br />

of two new reintroduction plantings. This funding will<br />

continue the momentum of successful reintroduction<br />

efforts. If we can successfully establish these two new<br />

populations and maintain the two large naturally-occurring<br />

sentry milk-vetch populations, Grand Canyon National<br />

Park will be well on its way to downlisting the<br />

species within the next ten years.<br />

Additional References<br />

Busco, J., E. Douglas, and J. Kapp. 2011. Preliminary<br />

pollination study on Sentry milk-vetch (Astragalus<br />

cremnophylax Barneby var. cremnophylax), Grand Canyon<br />

National Park’s only Endangered plant species.<br />

The <strong>Plant</strong> Press 35(1):11-12.<br />

Falk, M., J. Busco, L. Makarick, and A. Mathis.<br />

2011. The return of the “Watchman of the Gorge.” Endangered<br />

Species Bulletin, Summer 2011. Pp. 40-41.<br />

109


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

A Tale of Two Single Mountain Alpine Endemics:<br />

Packera franciscana and Erigeron mancus<br />

James F. Fowler, Carolyn Hull Sieg, Brian M. Casavant, and Addie E. Hite<br />

USFS Rocky Mountain Research Station, Flagstaff, AZ<br />

Abstract. Both the San Francisco Peaks ragwort, Packera franciscana and the La Sal daisy, Erigeron mancus are<br />

endemic to treeline/alpine habitats of the single mountain they inhabit. There is little habitat available for these plant<br />

species to migrate upward in a warming climate scenario. For P. franciscana, 2008 estimates indicate over 18,000<br />

ramets in a 4 m band along a recreational trail in the Arizona San Francisco Peaks, a trail-side population centroid of<br />

3667 m, and that the population is producing and dispersing seed. We also mapped the 2008 distribution of E.<br />

mancus patches along the La Sal Mountain crestline in <strong>Utah</strong>.<br />

Both the San Francisco Peaks ragwort, Packera franciscana<br />

(Greene) W.A. Weber and A. Löve, and the La<br />

Sal daisy, Erigeron mancus Rydberg, are endemic to<br />

treeline and alpine habitats of the single mountain they<br />

inhabit. Packera franciscana is known only from the<br />

San Francisco Peaks in Arizona (Greenman 1917, Trock<br />

2006) (Figure 1) where it has been reported to mostly<br />

occur between 3525 m and 3605 m elevation (Dexter<br />

2007) or more generally 3200-3800 m (Trock 2006)<br />

with a range size of 85 ha (Dexter 2007). Since the elevation<br />

of the highest peak on the mountain is 3851 m,<br />

there is little habitat available for the plant to migrate<br />

upward in a warming climate scenario, and it has been<br />

widely speculated that the species is vulnerable to extinction<br />

due to climate change. In 1985 the distribution<br />

of P. franciscana on the San Francisco Peaks was<br />

mapped (Dexter 2007), but prior to our study, no published<br />

data were available on species abundance. Erigeron<br />

mancus only inhabits the La Sal Mountains in <strong>Utah</strong><br />

(Cronquist 1947) (Figure 1) where it occurs in alpine<br />

meadows between 3000-3800 m elevation (Nesom<br />

2006). In sharp contrast to P. franciscana which predominately<br />

inhabits loose talus slopes (USFWS 1983),<br />

E. mancus occupies stable substrates, which greatly facilitates<br />

field measurements. No published information<br />

about the population biology of these species is available.<br />

Consequently, P. franciscana was listed as a<br />

Threatened species under the Endangered Species Act<br />

by the U.S. Fish and Wildlife Service (1983) and E.<br />

mancus is on the Forest Service Region Four Sensitive<br />

<strong>Plant</strong> List.<br />

Kruckeberg and Rabinowitz (1985) note that narrow<br />

endemics can be locally abundant in specific habitats<br />

but geographically restricted, a description that may fit<br />

both species. Biologists have noted that P. franciscana<br />

is fairly abundant locally (Trock 2006, USFWS 1983)<br />

and our observations concur. We know of no density or<br />

Figure 1. Map of the two study areas as isolated single<br />

mountains on the Colorado Plateau.<br />

population size data to support this observation, yet such<br />

data are critical for recovery of the species under the<br />

Endangered Species Act. In a changing climate scenario<br />

with increased temperatures and changes in<br />

amount, type, and patterns of precipitation, it becomes<br />

difficult to predict population trends. Our study will<br />

110


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

define baseline population densities along permanent<br />

transects under the current climate and allow the detection<br />

of future population trends. Specifically, our objectives<br />

are to: 1) establish a statistically robust sampling<br />

protocol for long-term population density trends; 2) determine<br />

the elevation of patch centroids along these<br />

transects to allow early detection of altitudinal migration<br />

driven by climate change; and 3) provide data for ongoing<br />

formal species assessments, management responses,<br />

and, in the case of P. franciscana, revision of the 20-<br />

year old Species Recovery Plan (Phillips and Phillips<br />

1987).<br />

METHODS<br />

In September 2008 (after the monsoon rains), we<br />

established an elevational transect along a designated<br />

recreational trail through Packera franciscana habitat to<br />

estimate the density of P. franciscana ramets, mid-September<br />

flowering/fruiting phenology, and the population<br />

centroid elevation as it intersects the trail (Figure 2).<br />

Sample points were established at 25 m intervals along a<br />

transect starting at 3550 m elevation and extending 1425<br />

m along the trail to an elevation of 3798 m. At each<br />

sample point we counted P. franciscana ramets (upright<br />

stems) within 12 individual 1 m 2 frames arranged to<br />

Figure 2. Location of the Packera franciscana trailside transect on the outslope of the volcanic caldera at and above<br />

treeline.<br />

111


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

allow flexibility for trail curvature (Figure 3). Sampling<br />

frames were omitted when they overlapped previously<br />

counted frames, covered recent trail maintenance areas,<br />

or covered vertical drop-off > 5 m. Counts of ramets<br />

with flower, fruit, or both were also made within each<br />

frame. Coordinates for latitude, longitude, and elevation<br />

were made for each sample point with a Trimble®<br />

Geo XT 2005 Series GPS. Descriptive statistics were<br />

calculated with SAS/STAT 9.1 (2002-2003). Population<br />

centroid was calculated as the mean elevation of<br />

occurrence weighted by the number of ramets / sample<br />

point.<br />

In July 2008, we mapped polygons of E. mancus<br />

patches with the Trimble® Geo XT 2005 Series GPS in<br />

three areas near and within Mt. Peale Research Natural<br />

Area, which is located in the Middle Group of peaks on<br />

La Sal Mountain. These polygons were plotted on a<br />

topographic map with ArcMap 9.2.<br />

RESULTS<br />

The September 2008 density estimate for P. franciscana<br />

along the recreational trail was 3.19 ramets / m 2<br />

ramets in the 4 m band along the transect. A population<br />

centroid was located at 3667 m elevation. We counted a<br />

total of 1881 ramets of which 91 percent were vegetative,<br />

eight percent were in fruit, and one percent were<br />

flowering. Only seven ramets were in both flower and<br />

fruit.<br />

Figure 3. Arrangement and sampling sequence of 1 m 2<br />

frames to measure ramet density.<br />

Erigeron mancus mapping work in July 2008 revealed<br />

a relatively continuous series of E. mancus<br />

patches along the west ridge up to Mt. Laurel in the La<br />

Sal Middle Group of peaks, from the talus field at 3725<br />

m down to 3475 m just above treeline, as well as along<br />

the La Sal Middle Group crestline at 3650 m (Figure 4).<br />

Our observations indicate that it can be abundant within<br />

its microhabitat niche on dry, windy ridgelines but less<br />

abundant to absent on nearby more mesic midslopes<br />

DISCUSSION<br />

Phillips and Peterson (1980) reported a P. franciscana<br />

population density range of 50-370 plants per 100<br />

m 2 on the San Francisco Peaks but did not clearly define<br />

plants as ramets or genets (clumps or clones) (Figure 5).<br />

However, later references to clump size would indicate<br />

that they were using the latter concept. On a per 100 m 2<br />

basis, our density measurements are similar at the upper<br />

end of their density range (319 vs. 370), which is probably<br />

a reflection of the different “plant” definitions.<br />

Given the difficulty of defining and counting clumps<br />

and clones in the field, ramets provide a more accurate<br />

way to assess population density. Even though ramet<br />

density may inflate the number of functional plants, it is<br />

an accurate reflection of photosynthetic and reproductive<br />

potential. Phillips and Peterson (1980) also reported<br />

that 13% of the P. franciscana plants were adult<br />

(sexually reproducing) which again is comparable to the<br />

9% of ramets we sampled which were flowering and/or<br />

fruiting. These results and our estimate of >18,000 P.<br />

franciscana ramets in a very small portion of its range<br />

would indicate that the species is persisting and reproducing.<br />

We interpret the successful production of fruit, which<br />

we observed actively dispersing by upslope winds in<br />

mid-September, as an indication that P. franciscana can<br />

sexually reproduce on the San Francisco Peaks. Seed<br />

viability studies may provide additional support for this<br />

interpretation. Examination of plant root systems would<br />

be necessary to determine if ramets originate from seed<br />

or from existing perennial rhizotamous clones. Rhizomes<br />

can produce large patches of ramets which may<br />

be the primary method of reproduction (USFWS 1983)<br />

but we also found single isolated ramets during our sampling<br />

which could be the result of seed dispersal or rhizome<br />

fragments moving downslope in the talus substrate<br />

P. franciscana inhabits. <strong>Plant</strong>s inhabiting the upper<br />

portions of talus slopes would seem to be the result<br />

of seed dispersal since avalanches and downslope creep<br />

of talus fields would carry existing P. franciscana plants<br />

downslope. We noted dead P. franciscana plants at the<br />

base of some avalanche chutes. The population centroid<br />

of 3667 m we measured is above the 3525-3605 m elevation<br />

range for most P. franciscana noted by Dexter<br />

(2007) and near the upper end of the 3350-3750 m main<br />

112


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 4. Erigeron mancus<br />

patches along the ridge up to<br />

Laurel Mt. and along the crestline<br />

of the Middle Group of La<br />

Sal mountain.<br />

occurrence range in earlier reports (Phillips and Peterson<br />

1980, U.S. Fish and Wildlife Service 1983). However,<br />

our transect is located on a drier west-southwest<br />

slope which may account for the higher occurrence elevation.<br />

More mesic slopes may have lower patch centroids;<br />

a hypothesis we intend to test by establishing a<br />

northeast aspect, trail-side transect in 2009.<br />

We plan the second trail-side transect and annual<br />

measurements of both transects to detect P. franciscana<br />

population trends. Sampling in subsequent years may<br />

indicate trends in population density, changes in September<br />

phenology, or elevational migration within its<br />

habitat. We also plan to measure the change in E.<br />

mancus density along an elevational transect through the<br />

E. mancus patches shown in Figure 4. By measuring<br />

patch widths along this elevational transect, we can cal-<br />

culate patch size and, using our density measurements,<br />

can then estimate population size for this area. Changes<br />

in population density and the elevation of population<br />

centroids over time for both species may allow detection<br />

of climate change effects as well as provide managers<br />

with accurate data on which to base land and recreation<br />

use decisions.<br />

ACKNOWLEDGEMENTS<br />

Thanks to Amanda Kuenzi and Suzy Neal for help<br />

with fieldwork and to the Coconino and Manti-La Sal<br />

National Forests for funding support. Thanks also to<br />

Shaula Hedwall and Barb Phillips who reviewed the<br />

previous version of this manuscript and provided access<br />

to internal reports.<br />

113


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 5. Clonal habit of Packera franciscana.<br />

LITERATURE CITED<br />

Cronquist, A. 1947. Revision of the North American<br />

species of Erigeron, north of Mexico. Brittonia 6(2):<br />

121-302.<br />

Dexter, L.R. 2007. Mapping impacts related to the<br />

Senecio franciscanus Greene. Final Report FS Agreement<br />

<strong>Number</strong>: 06-CR-11030416-777.<br />

Greenman, J.M.. 1917. Monograph of the North and<br />

Central American species of the Genus Senecio, Part II.<br />

Annals of the Missouri Botanical Garden. 4: 15-36.<br />

Kruckeberg, A.R. and D. Rabinowitz. 1985. Biological<br />

aspects of endemism in higher plants. Annual<br />

Reviews in Ecology and Systematics. 16: 447-479.<br />

Nesom, G.L. 2006. Erigeron. In: Flora of North<br />

America north of Mexico, Volume 20, Magnoliophyta:<br />

Asteridae, part 7: Asteraceae, part 2. Oxford University<br />

Press: 256-348.<br />

Phillips, A.M. III and E.M. Peterson. 1980. Status<br />

report: Senecio franciscanus. Office of Endangered<br />

Species, U.S. Fish and Wildlife Service, Albuquerque,<br />

NM. 13 p.<br />

Phillips, B.G. and A.M. Phillips III. 1987. San Francisco<br />

groundsel (Senecio franciscanus) recovery plan.<br />

U.S. Fish and Wildlife Service, Albuquerque, NM.<br />

SAS/STAT. 2002-2003. SAS/STAT® software, version<br />

9.1 of the SAS System for Windows 2000. SAS<br />

Institute Inc., Cary, NC.<br />

Trock, D.K. 2006. Packera. In: Flora of North<br />

America north of Mexico, Volume 20, Magnoliophyta:<br />

Asteridae, part 7: Asteraceae, part 2. Oxford University<br />

Press: 570-602.<br />

United States Fish and Wildlife Service. 1983. Endangered<br />

and threatened wildlife and plants; final rule to<br />

determine Senecio franciscanus (San Francisco Peaks<br />

groundsel) to be a threatened species and determination<br />

of its critical habitat. Federal Register 48: 52743-<br />

52747.<br />

Addendum<br />

The planned population size and density estimates<br />

for E. mancus were completed in summer 2009 and<br />

published in 2010 (Fowler and Smith 2010). We also<br />

added the second trailside transect for P. franciscana in<br />

2009 and published the 2008-2009 results in Fowler and<br />

Sieg (2010). A second P. franciscana manuscript covering<br />

the 2010-<strong>2012</strong> time frame is in preparation.<br />

Fowler, J.F. and C.H. Sieg. 2010. Density and elevational<br />

distribution of the San Francisco Peaks ragwort,<br />

Packera franciscana (Asteraceae), a threatened<br />

single-mountain endemic. Madroño 57(4):213-219.<br />

Fowler, J.F. and B. Smith. 2010. Erigeron mancus<br />

(Asteraceae) density as a baseline to detect future climate<br />

change in La Sal Mountain habitats. Journal Botanical<br />

Research Institute Texas 4(2):747-753.<br />

114


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Long-term Population Demographics and <strong>Plant</strong> Community Interactions<br />

of Penstemon harringtonii, an Endemic Species<br />

of Colorado’s Western Slope<br />

Thomas A. Grant III<br />

Program in Ecology, Colorado State University, Fort Collins, CO<br />

Michelle E. DePrenger-Levin,<br />

Denver Botanic Gardens Research and Conservation Dept, Denver, CO<br />

and Carol Dawson<br />

Bureau of Land Management Colorado State Office, Lakewood, CO<br />

Abstract. Penstemon harringtonii is an endemic species of Colorado’s western slope. Known from only six counties,<br />

Harrington’s penstemon is threatened primarily by habitat degradation and destruction in rural areas that are experiencing<br />

relatively rapid development and recreational pressure. Annual demographic monitoring since 1996 has<br />

not identified statistically significant changes in the overall number of rosettes, although significant inter-annual<br />

variation occurs at the two study sites. Additional research has focused upon the interactions of P. harringtonii with<br />

Artemisia tridentata (big sagebrush) and local plant species richness, and competition for soil moisture. Weak negative<br />

correlations between P. harringtonii and A. tridentata have been documented at both study sites, although the<br />

two sites have opposite trends in the correlation of P. harringtonii and species richness. Ordination techniques (nonmetric<br />

multi-dimensional scaling, NMS) are being explored as a means to find patterns that could increase our understanding<br />

of the rare species’ interactions with the dominant shrub (A. tridentata), local plant species diversity, and<br />

soil moisture. NMS found a positive correlation between species richness and the higher density P. harringtonii<br />

quadrats, although no strong relationships were identified between the rare species and soil moisture. Additional sites<br />

will be sampled in 2009 and 2010 to test hypotheses concerning the potential drivers of P. harringtonii density and<br />

provide guidance in the development of appropriate management and restoration methods.<br />

Long-term monitoring of Penstemon harringtonii<br />

Penland (Scrophulariaceae) was initiated in 1996 to determine<br />

population trends of this rare species. The penstemon<br />

is threatened by land development for homes<br />

and ski areas, oil and natural gas development, overgrazing,<br />

and off-road vehicle use (CoNPS 1997, Panjabi<br />

and Anderson 2006). Formerly a Category 2 candidate<br />

for listing under the Endangered Species Act (ESA), P.<br />

harringtonii is endemic to western Colorado (USDA<br />

NRCS 2009) and populations have been documented<br />

from six counties within the state (Eagle, Garfield,<br />

Grand, Pitkin, Routt, and Summit)(Panjabi and Anderson<br />

2006, Spackman et al. 1997). Currently, P. harringtonii<br />

is listed as a sensitive species by the U.S. Bureau<br />

of Land Management Colorado State Office and the<br />

USDA Forest Service Region 2. The species is ranked<br />

G3/S3 by The Nature Conservancy Natural Heritage<br />

ranking system (Spackman et al. 1997). Based upon the<br />

S3 ranking, the species is considered vulnerable to extirpation<br />

in the state and only 74 occurrences have been<br />

documented (Panjabi and Anderson 2006).<br />

The majority of populations occur in habitats dominated<br />

by Artemisia tridentata (big sagebrush), a hyd-<br />

raulic lifting species (Caldwell et al. 1998). The interaction<br />

between A. tridentata and growth of herbaceous<br />

plants such as P. harringtonii is unclear, although soil<br />

moisture and precipitation are acknowledged as limiting<br />

factors for primary production in semiarid sagebrush<br />

steppe ecosystems (Horton and Hart 1998).<br />

Initial goals of the project were to document population<br />

trends at two study sites, although this has been<br />

expanded into understanding the species’ relationship to<br />

A. tridentata density, soil moisture, and plant community<br />

composition with the ultimate goal of improving<br />

our management of the species and its habitat. We hypothesized<br />

that areas with lower densities of A. tridentata<br />

will have higher densities of P. harringtonii and<br />

greater species richness due to reduced competition for<br />

water. Although destruction or degradation of P. harringtonii’s<br />

habitat is the greatest threat, an understanding<br />

of the species’ interactions with big sagebrush and<br />

soil moisture may augment management of extant populations<br />

and restoration of degraded areas. Additionally,<br />

an understanding of the relationships between sagebrush,<br />

soil moisture and species richness may assist our<br />

ability to manage for diverse ecosystems.<br />

115


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

METHODS<br />

<strong>Plant</strong> Species<br />

Penstemon harringtonii is a potentially long-lived<br />

perennial forb in the Scrophulariaceae. A distinctive<br />

characteristic of the species is the exsertion of the two<br />

lower stamens from the blue to pink/lavender corollas<br />

(Figure 1). Flower production and seedling recruitment<br />

are thought to be episodic and probably related to seasonal<br />

precipitation and available soil moisture (Panjabi<br />

and Anderson 2006). The species is found in open sagebrush<br />

(A. tridentata) and less commonly in pinyonjuniper<br />

plant communities between 1951 and 2865m<br />

elevation. All major threats are based on increased human<br />

use and development in the region for housing, ski<br />

areas, resource extraction, grazing, and recreation<br />

(CoNPS 1997, Spackman et al 1997).<br />

Study Sites<br />

Two geographically and ecologically diverse populations<br />

of P. harringtonii have been monitored since<br />

1996. The Eagle study site is a sagebrush-steppe community<br />

near the town of Eagle, CO and is at an elevation<br />

of 2100m (Buckner and Bunin 1992). The site was<br />

roller-chopped in the 1980s (BLM personal communication)<br />

to decrease shrub cover and promote graminoid<br />

forage for cattle grazing and has relatively low sagebrush<br />

cover (6.99%). The Gypsum study area is located<br />

near the town of Gypsum, CO at an elevation of 2200m<br />

(Buckner and Bunin 1992). Relative to the Eagle site,<br />

this area has much higher cover of sagebrush (25.57%)<br />

and lower densities of P. harringtonii. Sagebrush cover<br />

was determined using the line-intercept method on the<br />

aerial cover of the shrubs and consisted of ten 60m transects<br />

per site. Based on a two-sample t-test, the amount<br />

of sagebrush cover was significantly different between<br />

the two study sites (P < 0.0001, alpha = 0.05, n = 10).<br />

Long-term Demographic Study<br />

At each site a 40 x 60m macroplot was installed at a<br />

location containing P. harringtonii and 1 x 60m quadrats<br />

were sampled within the macroplot based on a<br />

stratified random sampling method. The goal of the<br />

monitoring was to be statistically capable of detecting a<br />

20% change in the populations of P. harringtonii and<br />

was designed with the analysis having a power of 95%<br />

with a 1% chance of making a false-change error (Type<br />

I). Sample size and power analyses were conducted on<br />

the two initial years of data (1996 and 1997) to determine<br />

the appropriate number of quadrats for each site<br />

(Elzinga et al. 1998). The eight quadrats sampled at the<br />

Eagle site and 12 at Gypsum were sufficient to meet the<br />

desired power of the study. Within each quadrat the following<br />

data were collected: x and y coordinates, number<br />

of rosettes, and presence or absence of flowers, fruits<br />

Figure 1. Penstemon harringtonii (Photo by Carol Dawson).<br />

and herbivory. The quadrats were censused annually in<br />

early to mid-June. Repeated measures Analysis of Variance<br />

(ANOVA) and paired t-tests were utilized to determine<br />

statistically significant differences in rosette numbers<br />

over time or between years, respectively. Data were<br />

log transformed for statistical analysis.<br />

<strong>Plant</strong> Community Analysis<br />

In the summer of 2005 additional sampling of the<br />

macroplots was conducted as part of a Denver Botanic<br />

Garden internship investigating the relationship between<br />

A. tridentata density, soil moisture and the rare penstemon.<br />

Species richness and the density of sagebrush and<br />

Harrington’s penstemon were determined for 4 x 2m<br />

plots randomly located within the existing macroplots at<br />

the Eagle and Gypsum study sites. Sample size and<br />

power analysis determined that 17 and 20 plots were<br />

necessary for the Eagle and Gypsum study areas, respectively<br />

(Elzinga et al. 1998). Sample size was estimated<br />

using the 2005 data and a confidence level of<br />

90%. Analysis of the plant community data was conducted<br />

using Non-metric Multi-dimensional Scaling<br />

(NMS), a non-parametric multivariate ordination technique<br />

capable of detecting and describing vegetation<br />

patterns between the sites and correlating this inform-<br />

116


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

ation to species richness, soil moisture, and sagebrush<br />

density. The NMS used a Sorenson distance matrix and<br />

the presence or absence of plant species for the primary<br />

data matrix. The secondary matrix consisted of species<br />

richness per plot, categorical values for the ratio of Penstemon<br />

to Artemisia densities, and a dummy variable<br />

designating within which site the data are associated.<br />

Multi-Response Permutation Procedures (MRPP) provided<br />

a non-parametric statistical method to test for differences<br />

between groups (study sites) by comparing the<br />

heterogeneity within groups against the probability of<br />

occurrence by random chance. This type of randomization<br />

or permutation test provides an ‘A’ statistic and a<br />

‘P’ value that is used to determine if the two study sites<br />

are statistically significant from each other based on<br />

species compositions. Both NMS and MRPP were conducted<br />

in PC-Ord version 5.0 software (MjM Software<br />

1999). Linear regression and correlation coefficients<br />

(R 2 ) were determined between penstemon and sagebrush<br />

densities, soil moisture, and species richness using Microsoft<br />

Excel.<br />

RESULTS<br />

Long-term Demographic Study<br />

The 13 years of monitoring documented statistically<br />

significant variation in the number of rosettes per quadrat<br />

using a repeated measures ANOVA (Between subjects:<br />

P


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 4. Non-metric Multi-dimensional Scaling (NMS) of the plant species composition at the two study sites (Eagle<br />

and Gypsum). Triangles represent quadrats and plus signs are plant species with species codes as labels. Red vectors<br />

represent correlation coefficients with R 2 > 0.20 and are directionally aligned to plant species to which the vector<br />

variables are positively correlated. The length of the vector represents the magnitude of the correlation.<br />

data from the primary matrix. The vector for ‘species<br />

richness’ is oriented towards the sample space primarily<br />

occupied by the Eagle plots and represents a positive<br />

correlation between the Eagle samples and plots with<br />

higher species richness. The ‘dominance’ vector represents<br />

a ratio of P. harringtonii and A. tridentata densities<br />

and documents the positive correlation of the higher<br />

penstemon ratios with the Eagle samples. ‘Site’ is a<br />

categorical or ‘dummy’ variable necessary to code the<br />

sites and distinctly segregates the two study sites. Ordination<br />

analysis of soil moisture data and densities of<br />

Harrington’s penstemon or sagebrush did not determine<br />

any significant relationships or strong correlations. Although<br />

simple linear regressions determined that the<br />

Gypsum site had a weak negative relationship (R 2 =<br />

0.15) between the penstemon and sagebrush densities.<br />

Using linear regression to compare species richness with<br />

penstemon or sagebrush density, the two study sites had<br />

weak R 2 values, but the general trends were always<br />

opposite between the sites. At the Eagle site, species<br />

richness was positively related to penstemon density (R 2<br />

= 0.0856) or sagebrush density (R 2 = 0.0857). Conversely,<br />

at the Gypsum location the densities of penstemon<br />

and sagebrush were negatively correlated to species<br />

richness (R 2 = 0.108 and R 2 = 0.125, respectively).<br />

DISCUSSION<br />

The monitoring data documents that the P. harringtonii<br />

populations at Eagle and Gypsum are stable over<br />

the 13 years of monitoring (Figures 2 and 3), although<br />

they fluctuated greatly and reached the lowest population<br />

sizes during the droughts of the early 2000s. Increasing<br />

drought severity or frequency could have negative<br />

consequences for this species, especially if its habitat<br />

becomes more fragmented by development or degraded<br />

due to overuse. These results cannot be extrapolated<br />

to all populations of Harrington’s penstemon, but<br />

provide quantitative data concerning the population<br />

118


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

trends and supports the idea that the species is reproductive<br />

(data not presented) and recruitment is probably<br />

occurring at these locations. This monitoring project<br />

provided minimal data to inquire into the interactions of<br />

the plant community and resource use within the ecosystem.<br />

The multivariate ordination (NMS) analysis<br />

found interesting relationships between species richness<br />

and penstemon density (Figure 4), although linear regressions<br />

had weak results. We hypothesized that water<br />

is a driving resource in the system and that higher sagebrush<br />

density indirectly reduces species richness and<br />

penstemon density due to its control over the soil moisture<br />

and possibly space and nutrients. Our current data<br />

cannot support this hypothesis, but the initial analysis of<br />

the plant community data found interesting and relevant<br />

interactions that could lead to a better understanding of<br />

the rare species’ population and community dynamics.<br />

Sampling of additional sites and increased measuring of<br />

soil moisture will assist in the development of theories<br />

that relate sagebrush and penstemon densities to the<br />

availability of water and the effects of inter-specific<br />

competition.<br />

The Eagle study site was roller-chopped in the 1980s<br />

to reduce sagebrush cover and improve the system for<br />

grazing. This disturbance may have reduced sagebrush’s<br />

control over soil moisture and provided an appropriate<br />

disturbance for increased recruitment and establishment<br />

of P. harringtonii. An understanding of this ecosystem’s<br />

plant community dynamics and competition for water<br />

may provide clues to improving habitat of the rare penstemon,<br />

especially if the rate of habitat loss continues to<br />

accelerate in this rapidly developing region. If we improve<br />

our knowledge of the resources (water) and processes<br />

(disturbance) that regulate P. harringtonii it may<br />

be possible to develop management and restoration<br />

practices that promote higher density populations of the<br />

species by the manipulation of sagebrush cover and surface<br />

disturbance.<br />

REFERENCES<br />

Buckner, D.L. and J.E. Bunin. 1992. Final report:<br />

1990/91 status report for Penstemon harringtonii<br />

Penland. Unpublished report prepared for the Colorado<br />

Natural Areas Program, Denver, CO by Esco Associates,<br />

Inc., Boulder, CO.<br />

Caldwell, M.M., T.E. Dawson, and J.H. Richards.<br />

1998. Hydraulic lift: Consequences of water efflux from<br />

the roots of plants. Oecologia 113:151-161.<br />

Colorado <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 1997. Rare <strong>Plant</strong>s of<br />

Colorado. 2nd Edition. Falcon Press. Helena, MT.<br />

Elzinga, C.L., D.W. Salzer and J.W. Willoughby.<br />

1998. Measuring and Monitoring <strong>Plant</strong> Populations.<br />

BLM Technical Reference 1730-1 (BLM/RS/ST-<br />

98/005+1730).<br />

Horton, J.L. and S.C. Hart. 1998. Hydraulic lift: a<br />

potentially important ecosystem process. TREE 13(6):<br />

232-235.<br />

Panjabi, S.S. and D.G. Anderson. (2006, June 30).<br />

Penstemon harringtonii Penland (Harrington’s beardtongue):<br />

a technical conservation assessment. [Online].<br />

USDA Forest Service, Rocky Mountain Region. Available:<br />

http://www.fs.fed.us/r2/projects/scp/assessments/<br />

penstemonharringtonii.pdf [1/7/09].<br />

PC-ORD Version 5.0. 2005. MjM Software, Gleneden<br />

Beach, OR, USA.<br />

Spackman, S., B. Jennings, J. Coles, C. Dawson, M.<br />

Minton, A. Kratz, and C. Spurrier. 1997. Colorado Rare<br />

<strong>Plant</strong> Field Guide. Prepared for the Bureau of Land<br />

Management, the U.S. Forest Service and the U.S. Fish<br />

and Wildlife Service by the Colorado Natural Heritage<br />

Program.<br />

USDA, NRCS. 2009. The PLANTS Database (http://<br />

plants.usda.gov, 27 May 2009). National <strong>Plant</strong> Data<br />

Center, Baton Rouge, LA 70874-4490 USA.<br />

119


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Conservation and Restoration Research<br />

at The Arboretum at Flagstaff<br />

Kristin E. Haskins and Sheila Murray<br />

The Arboretum at Flagstaff, Flagstaff, AZ<br />

Abstract. The Colorado Plateau is experiencing increased climate change effects and population expansion. Many<br />

native plant species are at risk for becoming rare or threatened, and it is challenging to secure local, native plant seed<br />

for use in restoration. Here we highlight two examples of our conservation efforts for Astragalus cremnophylax var.<br />

cremnophylax in the Grand Canyon and Purshia subintegra in the Verde Valley of central Arizona, and discuss our<br />

involvement in a local native plant propagation movement conducted by the US Forest Service and the Museum of<br />

Northern Arizona. The Arboretum has been working toward rare plant conservation and restoration efforts for over<br />

25 years.<br />

SENTRY MILK-VETCH CONSERVATION<br />

Astragalus cremnophylax var. cremnophylax (Figure<br />

1) was listed as a species of concern in 1980 and was<br />

bumped up to endangered species status in 1990. This<br />

tiny legume is found only at Grand Canyon National<br />

Park (GCNP) in limestone outcrops. Threats to this species<br />

include development of the park and climate<br />

change.<br />

In 2005, The Arboretum began working with GCNP<br />

and The U.S. Fish and Wildlife Service to conserve this<br />

rare species. The initial task was to increase seed availability<br />

through ex-situ propagation. Typically, seeds are<br />

propagated in sterile potting mix that can provide mixed<br />

results in terms of seed germination and growth. In an<br />

effort to improve seedling performance, we examined<br />

the effects of different soil treatments: 1) potting soil<br />

with Rhizobium added and 2) potting soil with a native<br />

Figure 2. Astragalus cremnophylax var. cremnophylax<br />

grown with a native soil inoculum (left) and in standard,<br />

sterile potting soil (right). Photo by Sheila Murray.<br />

inoculum added versus traditional potting soil. After<br />

five months of growth, seedlings propagated with a native<br />

soil inoculum had significantly greater aboveground<br />

volume than seedlings grown in either of the other two<br />

treatments (Figure 2). We are currently tracking these<br />

plants to determine if these differences will translate<br />

into increased seed production. Ultimately, we aim to<br />

conserve Sentry Milk-vetch by expanding established<br />

and adding new populations at GCNP.<br />

Figure 1. Astragalus cremnophylax var. cremnophylax<br />

in bloom. Photo by Julie Crawford.<br />

ARIZONA CLIFFROSE RESTORATION<br />

Purshia subintegra, or Arizona cliffrose, is known<br />

from four populations in central Arizona, with the largest<br />

population occurring in the Verde Valley. This xeric,<br />

evergreen member of the Rosaceae was listed as endangered<br />

in 1984. Major threats include development, overgrazing<br />

and climate change. The Arizona Department of<br />

Transportation funded work by The Arboretum from<br />

120


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

1996-2000 in response to road construction that would<br />

result in a loss of Arizona cliffrose habitat. The Arboretum<br />

developed a protocol for propagating Arizona cliffrose<br />

via cuttings (Figure 3), as recent droughts had prevented<br />

the species from producing seed. Additionally,<br />

The Arboretum examined ways in which the propagated<br />

cuttings (Figure 4) could be put back in the field onto<br />

protected sites. Since out-planting in 2001, the Research<br />

Department has been involved in monitoring the new<br />

populations. We are happy to report that the new populations<br />

are doing well.<br />

ARIZONA NATIVE PLANT PROPAGATION<br />

In 2007, The Arboretum began a collaborative project<br />

with the U.S. Forest Service and The Museum of<br />

Northern Arizona to collect and propagate native seeds<br />

for use in local restoration efforts (Figure 5). This project<br />

arose in response to a high demand and lack of supply<br />

of local seed genotypes that were crucially needed<br />

after large scale forest fires hit the area in 2002.<br />

The first phase of the project is to collect and propagate<br />

seeds of native species that appeal to land managers<br />

for use in re-vegetation projects. <strong>Plant</strong> species are being<br />

chosen for their wildlife forage quality and likelihood of<br />

propagation success. We are also focusing on species<br />

that are not already in commercial production. Our goal<br />

is to start small, but eventually produce a reliable source<br />

for local seed genotypes that can be used by local land<br />

managers.<br />

ACKNOWLEDGEMENTS<br />

The Research Department at The Arboretum at Flagstaff<br />

(a.k.a. Kris and Sheila) would like to thank all of<br />

our wonderful volunteers and the following groups for<br />

financial support: The U.S. Fish and Wildlife Service,<br />

Arizona Department of Transportation, and The U.S.<br />

Forest Service.<br />

Figure 4. An Arboretum volunteer helps re-pot Purshia<br />

subintegra in the research greenhouse. Photo by K.<br />

Haskins.<br />

Figure 3. Sheila Murray collects cuttings of Purshia<br />

subintegra in the Verde Valley, AZ. Photo by Joyce<br />

Maschinski.<br />

Figure 5. The research greenhouse at The Arboretum;<br />

where plant propagation will take place. Photo by K.<br />

Haskins.<br />

121


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

The Digital Atlas of <strong>Utah</strong> <strong>Plant</strong>s:<br />

Determining Patterns of Biodiversity and Rarity<br />

Leila M. Shultz, R. Douglas Ramsey, Wanda Lindquist, and C. Garrard<br />

Floristics Lab and Remote Sensing/GIS Lab, <strong>Utah</strong> State University, Logan, UT<br />

Abstract. The digital Atlas of <strong>Utah</strong> <strong>Plant</strong>s is a web-based revision of the Atlas of the Vascular <strong>Plant</strong>s of <strong>Utah</strong> by B.<br />

Albee, L. Shultz, and S. Goodrich, published in 1988 by the University of <strong>Utah</strong> Museum of Natural History. The hard<br />

copy version provided distribution maps for 2,438 native or naturalized plant species and an appendix with generalized<br />

locations for 399 additional species that are either extremely rare, recently introduced, or at the edges of their<br />

ranges – and known from one population. The new digitized version allows analysis of original data from pre-1988<br />

herbarium collections (mapped at coarse level, resolution at approximately 10 km 2 ) and brings in herbarium records<br />

for post-1988 collections (most of which are mapped at high resolution with global positioning devices) as well as<br />

more than 6,000 observations and records of rare species from the database for the <strong>Utah</strong> Natural Heritage program.<br />

While specific location sites for rare species are not displayed on the web-based version, the presence of rare species<br />

is highlighted on the grid map displayed for each species. The digital version provides a tool for tracking reports of<br />

new records as well as a tool for analyzing patterns of diversity. Open access to these records is currently available<br />

and species lists have been compiled for each major ecoregion in the state.<br />

The on-line version of the Atlas of <strong>Utah</strong> <strong>Plant</strong>s by<br />

Leila Shultz, R. Douglas Ramsey, and Wanda Lindquist<br />

is a revision of the Atlas of the Vascular <strong>Plant</strong>s of <strong>Utah</strong><br />

(Albee et al.1988). The new digital version displays a<br />

source code for each mapped point, new records, general<br />

locations for rare species, and nomenclatural updates.<br />

The original Atlas was based on the authors’ examination<br />

of approximately 400,000 herbarium specimens<br />

of <strong>Utah</strong> plants housed within the natural history collections<br />

of Brigham Young University (BRY), University<br />

of <strong>Utah</strong> (UT), <strong>Utah</strong> State University (UTC), the Forest<br />

Service Herbarium in Ogden (OGDF), and several Bureau<br />

of Land Management and Park Service herbaria.<br />

Although the original maps were hand-plotted on a gridded<br />

base map from specimens that were not entered in a<br />

database, the herbarium location for each voucher was<br />

color-coded on the map. These original maps are archived<br />

at the University of <strong>Utah</strong> Museum of Natural<br />

History’s Garrett Herbarium (Shultz et al. 1998; Ramsey<br />

and Shultz 2004).<br />

METHODS<br />

Technicians at the Remote Sensing/Geographic Information<br />

System laboratory of <strong>Utah</strong> State University<br />

hand-digitized the original maps. The first on-line version<br />

was sorted by family name with access through a<br />

web site. Development of GIS layers follows the methodology<br />

reported in Ramsey and Shultz (2004). The<br />

1992 version remained unchanged until the development<br />

of the digital version (Shultz et al. 2007).<br />

Voucher specimens on which the atlas is based represent<br />

more than 150 years of work by scores of dedicated<br />

professionals and amateurs, many of whom devoted<br />

their lives to tracking down unusual plants in some of<br />

the most isolated and physically challenging areas in<br />

North America. Authors of the original atlas (Beverly<br />

Albee, Leila Shultz, and Sherel Goodrich) spent approximately<br />

seven years critically examining and mapping<br />

locations for these approximately 400,000 collections<br />

that are housed primarily in <strong>Utah</strong> herbaria. Each<br />

point on the original maps is color-coded to show which<br />

herbarium houses the voucher for a particular record.<br />

Original maps are archived at the University of <strong>Utah</strong>’s<br />

Garrett Herbarium (UT). While the lack of specific<br />

voucher data is somewhat of a handicap, this problem<br />

will be corrected as on-line retrieval systems are developed<br />

through the developing consortium of herbaria in<br />

<strong>Utah</strong>. Once these data are available, users should be able<br />

to retrieve specific location records through the internet<br />

links to individual herbaria, or a consortium of herbaria<br />

(see the Intermountain Herbarium website for new reports,<br />

at http://herbarium.usu.edu/holdings_ specimen _<br />

database.htm).<br />

RESULTS<br />

Digitized version. The digital version allows for<br />

analysis of distribution patterns and patterns of diversity<br />

within the state. The composite geographic information<br />

system for 2,840 plant species provides approximately<br />

77,000 locations (some with multiple records) in 10 X<br />

10 km grids (roughly equal to a township). The digital<br />

122


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

version displays new geographic records, new species,<br />

and the more than 500 nomenclatural changes that have<br />

taken place since publication of the hard copy in 1988<br />

(Figure 1). Additions include approximately 8,000 new<br />

location records (including 6,000 locality records from<br />

the <strong>Utah</strong> Natural Heritage program data files that are<br />

mapped at the resolution of township and range), and<br />

new entries for more than 400 rare species that were not<br />

included in the 1988 publication (including new reports<br />

in Welsh et al. 2003). It is based on a real-time mapping<br />

system that draws from numerous data layers and displays<br />

records on a map of <strong>Utah</strong> with a choice of backgrounds<br />

(satellite image, state map with county boundaries,<br />

or one with Nature Conservancy ecoregion boundaries).<br />

When viewing the distribution map, clicking on a<br />

point allows one to see the source of the information<br />

(see explanations below) as well as the elevation of the<br />

specific location.<br />

Nomenclatural updates and addition of rare species.<br />

Name changes resulting from nomenclatural revisions<br />

(Flora of North America 1993 – 2006) involve approximately<br />

16% of the names in the <strong>Utah</strong> flora. We worked<br />

to make the transition to new names as painless as possible:<br />

"old names" are retained, but shown in italics.<br />

When you click on an italicized name, you will be taken<br />

to the accepted "new name". Common names and nomenclatural<br />

changes are based on the USDA <strong>Plant</strong>s Database<br />

(2009). Rare plants are highlighted in red letters,<br />

with the species taken from the state sensitive plant list<br />

(based on the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> [2009] and Bureau<br />

of Land Management Sensitive Species List).<br />

Species checklists. Species checklists for each of the<br />

major ecoregions in the state are provided on the web<br />

page. The checklists are found by clicking on the name<br />

of an eco-region on the right side of the web page. Information<br />

for each species listed includes the common<br />

Figure 1. Sample Page from the Digital Atlas of <strong>Utah</strong> <strong>Plant</strong>s (http://earth.gis.usu.edu/plants) showing distribution of<br />

Abies concolor (White fir). A color-corrected satellite image is chosen as the background (note other choices available,<br />

including ecoregions), courtesy of the Remote Sensing/GIS Lab at <strong>Utah</strong> State University.<br />

123


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 2. Species richness patterns based on voucher specimens (Albee et al. 1988), relative to a hexagonal grid sampling<br />

frame (EPA 649 km 2 sampling frame). Areas highest in species richness are shown in red, with lowest species<br />

richness shown in blue. Patterns of species richness generally correlate with elevation: the higher the richness, the<br />

higher the elevation (Ramsey and Shultz 2004).<br />

name, family name, growth form (tree, shrub, grass, or<br />

forb), known elevational range within <strong>Utah</strong>, acronym<br />

code (USDA PLANTS database), and notation as to<br />

whether the species is native or introduced. Rare plants<br />

are also highlighted in red on these checklists.<br />

Analysis of biodiversity patterns. The digital version<br />

allows reports of total species richness by ecoregion as<br />

well as diversity within a uniform grid system (Figure<br />

2). The layers also can be manipulated to analyze patterns<br />

of species diversity between ecoregions. For comparisons<br />

of ecoregions within the state, see Shultz and<br />

others (2000).<br />

DISCUSSION<br />

Anyone attempting to represent the distribution of a<br />

biological organism knows that distributions are not<br />

124<br />

static and that maps can do no better than represent the<br />

distribution of a species at a specified point in time.<br />

Changing landscapes have a profound effect on the distribution<br />

of a species. In addition, the development of<br />

an atlas of plants contains a number of inherent problems<br />

-- primarily regarding accurate identification and<br />

scale. Due to the diversity of the vascular plants (with<br />

more than 20,000 species in North America according to<br />

the Flora of North America Editorial Committee 1992--<br />

2006), reports of plant species generally cannot be<br />

trusted unless accompanied by a voucher specimen.<br />

That constraint severely limits the sample size and<br />

skews the kind of species represented by vouchers. In<br />

general, common species are under-represented in herbarium<br />

collections. However, rare occurrences are usually<br />

well-represented in herbarium collections and con-


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

sequently, distribution maps of rare species are highly<br />

reliable as to the total range of a species.<br />

When mapping from herbarium vouchers, botanists<br />

have the advantage of having a verifiable report that can<br />

be re-examined (and re-mapped) if the distribution is<br />

questionable. There is much biological and climatically<br />

important information to be gained from the mapping of<br />

a species distribution, but the user of a map should understand<br />

how the data are collected in order to understand<br />

what kinds of analyses can be performed. The<br />

first thing a user should understand is that a dot on a<br />

map will not guide the user to a specific spot on the<br />

ground. Imprecise location data for older herbarium<br />

specimens makes it impossible to map most reported<br />

plant distributions at a fine scale. New herbarium records,<br />

however, generally provide location data collected<br />

by global positioning systems. While these readings<br />

might be off by a hundred meters or so, the level of<br />

accuracy represents a considerable improvement over<br />

the older records. In order to understand how to interpret<br />

the reported distributions, one must first understand<br />

that the maps represent ranges of species rather than<br />

precise locations on the landscape.<br />

In developing the first atlas for <strong>Utah</strong> plants (Albee et<br />

al. 1988), the authors were fully aware that development<br />

of a collection database would be preferable to simply<br />

creating dot maps. However, constraints of time, cost of<br />

equipment, and limits of available technology in the<br />

“early days” of computers made it impossible to consider<br />

the development of a collection database. By our<br />

rough estimate, we calculated that such an undertaking<br />

would take at least twenty years. In addition, we knew<br />

that we would be overwhelmed by the problems inherent<br />

in mapping from a literal translation of herbarium<br />

data. We chose instead to spend our time checking the<br />

identification of each specimen, using the most current<br />

monographic or floristic treatments available. If a collection<br />

could not be mapped to the accuracy of township<br />

and range, we did not represent it with a dot. We could<br />

not map between twenty and thirty percent of all collections,<br />

and specimens from locations that were already<br />

mapped were not mapped again. We did, however,<br />

color-code each dot as to the herbarium from which the<br />

record was obtained. A questionable distribution, or<br />

one of particular interest, can thus be traced by consulting<br />

the archived maps at the University of <strong>Utah</strong>. A correction<br />

of distribution maps thus requires re-examination<br />

of specimens – a procedure that is highly recommended<br />

in the event of new studies or generic revisions.<br />

The lack of database-generated maps is not a great<br />

handicap at this time, but there should be considerable<br />

improvement in the future. If we (the authors) had access<br />

to a graphics tablet or a system for creating bar<br />

codes when we initiated the project, we would have<br />

simply placed a bar code on the specimen and linked it<br />

to a dot on a base map underlain by a graphics tablet.<br />

That kind of system would have been time-efficient,<br />

allowing later linkage to a database. Undoubtedly, new<br />

maps will be generated as collection databases are developed,<br />

and we can only hope that they will represent<br />

greater scale as well as a better way to track species distributions<br />

through time.<br />

For the present, the digitized atlas provides good estimates<br />

of species ranges within <strong>Utah</strong>, a mechanism for<br />

generating species lists for any specified area, relatively<br />

current nomenclature, and highly accurate estimates as<br />

to the number and distribution of rare species in the<br />

state. The authors of this paper encourage the use of the<br />

digitized atlas and invite readers to visit the “Virtual<br />

<strong>Utah</strong>” website at http://earth.gis.usu.edu/utah/.<br />

ACKNOWLEDGEMENTS<br />

A grant from the Bureau of Land Management made<br />

this modern (post-2005) digital revision possible. Curators<br />

of the S.L. Welsh Herbarium of Brigham Young<br />

University (BRY), the Garrett Herbarium of University<br />

of <strong>Utah</strong> (UT), and the Intermountain Herbarium of <strong>Utah</strong><br />

State University (UTC), provided enormous support<br />

throughout the years of examination of specimens.<br />

Since 2004, digital records from UTC and <strong>Utah</strong> Valley<br />

University (UVSC) have been added, for which we<br />

thank Michael Piep and Renee Van Buren. Staff associated<br />

with collections housed with various National<br />

Parks, Forests, and Bureau of Land Management offices<br />

in the state helped by providing access to collections. R.<br />

Douglas Ramsey had the vision to see the potential in<br />

developing the geospatial format and provided the technical<br />

support that made the project happen. Ben Franklin<br />

of the Division of Wildlife Resources sent records of<br />

rare species collections and observations in buffered,<br />

digitized format--a contribution of inestimable value.<br />

Walt Fertig provided his voucher data for Grand Staircase<br />

Escalante Monument, specimens deposited at BRY<br />

and UTC. Bonnie Banner, Thad Tilton, Tom Van Neil,<br />

Kent Braddy, and Chris Garrard of the Remote Sensing<br />

Lab of USU helped develop the geo-referenced database<br />

and create the original digital interface. Wanda<br />

Lindquist provided the programming expertise that allowed<br />

us to incorporate nomenclatural revisions, link to<br />

new data layers, and create species checklists. She designed<br />

the new web interface and integrated the complex<br />

system of new data layers.<br />

LITERATURE CITED<br />

Albee, B.A., L.M. Shultz, and S. Goodrich. 1988.<br />

Atlas of the Vascular <strong>Plant</strong>s of <strong>Utah</strong>. Univ. of <strong>Utah</strong> Museum<br />

of Nat. History, Occasional Publ. no.7. 670 p.<br />

Flora of North America Editorial Committee,<br />

eds. 1993+. Flora of North America North of Mexico.<br />

12+ vols. New York and Oxford. (Vol. 1, 1993;<br />

125


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

vol. 2, 1993; vol. 3, 1997; vol. 4, 2003; vol. 5, 2005;<br />

vol. 19, 2006; vol. 20, 2006; vol. 21, 2006; vol. 22,<br />

2000; vol. 23, 2002; vol. 25, 2003; vol. 26, 2002.)<br />

Ramsey, R.D. and L. Shultz. 2004. Evaluating the<br />

Geographic Distribution of <strong>Plant</strong>s in <strong>Utah</strong> from the Vascular<br />

Atlas of <strong>Utah</strong> <strong>Plant</strong>s. W. N. Amer. Nat. 64: 421-<br />

432.<br />

Shultz, L.M., N. Morin, and R.D. Ramsey. 1998.<br />

Floristics in North America: Tracking Rare Species<br />

Electronically. In C.I. Peng & P. P. Lowrey II, eds., Institute<br />

of Botany, Academia Sinica Monograph 16:<br />

259—273.<br />

Shultz, L.M., R.D. Ramsey, W. Lindquist. 2007.<br />

Digital Atlas of <strong>Utah</strong> <strong>Plant</strong>s: http://earth.gis.usu.edu/<br />

plants [revision ongoing, when citing, please give date<br />

of access]<br />

Shultz, L.M., R.D. Ramsey, and P. Terletzky. 2000.<br />

Using an atlas of vascular plants to determine patterns<br />

of biodiversity in <strong>Utah</strong>. Ecological <strong>Society</strong> of America<br />

85 th Annual Meetings August 6—10. ESA Abstracts:<br />

337.<br />

USDA, NRCS. 2009. The PLANTS Database (http://<br />

plants.usda.gov, Accessed May 2007). National <strong>Plant</strong><br />

Data Center, Baton Rouge, LA 70874-4490 USA.<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 2009. <strong>Utah</strong> Rare <strong>Plant</strong><br />

Guide (http://www.utahrareplants.org/rpg_species.<br />

html#All)<br />

Welsh, S.L., N.D. Atwood, S. Goodrich, and L.C.<br />

Higgins. 2003. A <strong>Utah</strong> Flora, third edition. Brigham<br />

Young University, Provo, UT.<br />

Addendum<br />

A tool that allows users to extract species lists in database form, with accompanying information about species,<br />

has been added to the Digital Database website. Users can draw a polygon of any size within the <strong>Utah</strong> borders (using<br />

the ‘limit area’ command) and download a species list in .dbf format (using the ‘get list’ command). The list will be<br />

accompanied by information that includes common names, currently accepted acronyms, growth form, elevational<br />

range, nativity, etc. in an electronic format that can be incorporated into spreadsheets or database programs. The<br />

reference for the site is:<br />

Shultz, L.M., R.D. Ramsey, W. Lindquist, and C. Garrard. <strong>Utah</strong> State University, Logan, UT. (http://earth.gis.usu.<br />

edu/plants/).<br />

126


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Molecular Genetic Diversity and Differentiation in Clay Phacelia<br />

(Phacelia argillacea Atwood: Hydrophyllaceae)<br />

Steven Harrison<br />

Brigham Young University, Provo, UT<br />

Susan E. Meyer,<br />

US Forest Service Rocky Mountain Research Station, Shrub Sciences Laboratory, Provo UT<br />

and Mikel Stevens,<br />

Brigham Young University, Provo, UT<br />

Abstract. Clay phacelia (Phacelia argillacea Atwood) was listed as federally endangered in 1978. It is known from<br />

only two populations in Spanish Fork Canyon, <strong>Utah</strong>. Samples were taken in each of three years from each of these<br />

two populations. We used AFLP markers to assess the genetic relatedness between the two populations and degree of<br />

differentiation between P. argillacea and three of its congeners. Six AFLP primer combinations resulted in 535 reliable<br />

marker loci of which 124 were polymorphic. Phacelia argillacea is genetically distinct from both its close and<br />

distant congeners. The two P. argillacea populations were not strongly differentiated, suggesting that gene flow between<br />

these populations probably occurred historically. In contrast, cohorts establishing in different years within a<br />

population were often genetically differentiated. Sampling in a single year would seriously underestimate genetic<br />

diversity in this species.<br />

Phacelia argillacea Atwood (clay phacelia) is a narrow<br />

endemic presently known from two locations approximately<br />

8 kilometers apart in the Spanish Fork Canyon,<br />

<strong>Utah</strong> County, <strong>Utah</strong> (Figure 1). The genus Phacelia<br />

is the largest in the Hydrophyllaceae. Phacelia argillacea<br />

is a member of the Crenulatae group of section Phacelia,<br />

subgenus Phacelia and is thought to be most<br />

closely allied to P. glandulosa Nutt., a species of wide<br />

distribution in extreme eastern <strong>Utah</strong>, western Colorado,<br />

Wyoming, eastern Idaho, and Montana (Atwood 1975).<br />

Another apparently closely allied species is the recently<br />

described P. argylensis Atwood (Welsh et al. 2003),<br />

known only from the type location in Argyle Canyon,<br />

Carbon County, <strong>Utah</strong> (Figure 1). The Crenulatae group<br />

(which consists of approximately 35 species) differs<br />

from other species of Phacelia in producing four-seeded<br />

capsules, faveolate seeds with a central ridge on the<br />

ventral side, and have a chromosome number of n=11<br />

(Atwood 1975).<br />

Recent taxonomic work in the genus Phacelia has<br />

confirmed the placement of P. glandulosa within the<br />

Crenulatae but has not included either P. argillacea or<br />

P. argylensis (Garrison 2007, Gilbert et al. 2005). Our<br />

goal was to examine molecular genetic diversity within<br />

and among the two known populations of P. argillacea<br />

as a necessary step in designing strategies for introducing<br />

new populations of this species on public land. We<br />

also wanted to make a preliminary assessment of the<br />

degree of differentiation between P. argillacea and its<br />

close congeners P. glandulosa and P. argylensis. We<br />

included the more distant congener P. crenulata Torr.<br />

ex S. Wats. as an out group in the analysis.<br />

Phacelia argillacea is an annual or biennial and has<br />

years when few or no actively growing plants are present.<br />

Its seeds germinate from spring to late summer or<br />

early fall and produce a rosette of leaves which grows<br />

during the winter months and bolts in the spring to produce<br />

a flowering shoot (Armstrong 1992, Meyer personal<br />

observation). The species is an edaphic endemic<br />

that is confined to steep hillsides of the Green River<br />

shale formation. Because of its restricted habitat and<br />

small and widely fluctuating population size, P. argillacea<br />

was declared an endangered species in 1978 (US<br />

Fish and Wildlife Service 1978, 1989).<br />

In 1990, The Nature Conservancy purchased the<br />

Tucker site, which at the time was the only known extant<br />

population for P. argillacea, and fenced it to prevent<br />

damage from grazing and trampling by deer and<br />

sheep and from disturbance caused by highway and railroad<br />

construction (Armstrong 1992). The plant was<br />

later rediscovered at the Railroad site, further down the<br />

canyon (Figure 1, inset). This population is on private<br />

land and is not fenced or managed for conservation.<br />

Known impacts to the Railroad site are highway widening<br />

coupled with the construction of a retaining wall,<br />

subsequent erosion, and trailing and grazing of domestic<br />

and native ungulates. P. argillacea may also be threatened<br />

by invasive weeds and drought.<br />

127


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 1. Collection sites for four species of Phacelia included in the AFLP (Amplified Fragment Length Polymorphism)<br />

study. Tissue samples were field-collected or greenhouse-grown tissue for P. argillacea and P. glandulosa<br />

for Site 2, while samples from the other P. glandulosa sites and for P. argylensis were obtained from specimens in<br />

the Brigham Young University herbarium. (Inset shows Spanish Fork Canyon, <strong>Utah</strong>, with extant P. argillacea subpopulations<br />

in black, reintroduction sites in green).<br />

Through U.S. Fish and Wildlife funding, a working<br />

group was organized in 2004 to focus on the introduction<br />

of P. argillacea into suitable and presumably previously<br />

occupied habitat on public land in Spanish Fork<br />

Canyon, an action recommended in the recovery plan<br />

for this species (US Fish and Wildlife Service 1989).<br />

Because many endangered plants with small population<br />

sizes and fragmented populations, such as P. argillacea,<br />

suffer higher risk of extinction due to genetic drift and<br />

inbreeding as well as stochastic environmental effects,<br />

an introduction program is almost essential for rare<br />

plants like P. argillacea (Kang et al. 2005). In 2007,<br />

seeds produced from greenhouse-grown individuals<br />

from the Tucker site were introduced at two new sites<br />

on US Forest Service land. The inset in Figure 1 depicts<br />

the location of the reintroduction sites (Mill Fork and<br />

Tie Fork) with reference to the extant P. argillacea<br />

populations.<br />

Analysis of the genetic diversity within and between<br />

the P. argillacea populations was also necessary because<br />

the species had never been studied at the molecular<br />

level. Analysis of the genetic diversity of an endan-<br />

gered plant species is a key element in the estimation of<br />

the viability of a population and can assist in conservation<br />

programs (Ronikier 2002, Kang et al. 2005). Data<br />

were also needed to address the question of whether P.<br />

argillacea was truly a distinct species or simply a disjunct<br />

population of P. glandulosa. Amplified Fragment<br />

Length Polymorphism (AFLP) was the molecular<br />

marker chosen to quantify genetic diversity. AFLPs<br />

were chosen because they have the potential to resolve<br />

genetic differences for individual identification and require<br />

no prior sequence knowledge of the organism. In<br />

this study AFLP analysis was used to address the following<br />

questions with regard to reintroduction of P. argillacea:<br />

(a) What is the level of genetic diversity in the<br />

two populations? (b) Is there genetic differentiation between<br />

the two populations? (c) Do samples collected<br />

within a population within a single year represent a random<br />

sample of genetic variation, or is there genetic differentiation<br />

between years? (d) Is P. argillacea genetically<br />

distinct from its close congeners P. argylensis and<br />

P. glandulosa? The answers to these questions should<br />

help to inform reintroduction efforts for this organism.<br />

128


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

MATERIALS AND METHODS<br />

Sample Collection<br />

Phacelia argillacea leaf tissue samples were collected<br />

from Tucker in the 2006 and 2008 field seasons<br />

and from Railroad in the 2006, 2007, and 2008 field<br />

seasons. A single basal leaf was removed nondestructively<br />

from each individual. The 2004 Tucker samples<br />

represent half-sibling progeny from wild-collected seeds<br />

of 15 maternal individuals (a total of 53 plants) in the<br />

2004 field season. These individuals were grown for<br />

seed production. Phacelia crenulata samples represented<br />

individuals greenhouse-grown from seeds of a<br />

bulk wild collection. Phacelia argylensis and some P.<br />

glandulosa samples were collected from Brigham<br />

Young University Herbarium; specimens were annotated<br />

as sampled for this study. The remaining P. glandulosa<br />

samples represent bulk population samples from<br />

two closely adjacent populations collected by Frank<br />

Smith in 2007 (Figure 1).<br />

DNA Extraction and AFLP Analysis<br />

Fresh leaf tissue samples for DNA extraction were<br />

dried over silica gel, lypophilized, or frozen at -80C immediately<br />

after collection. DNA was extracted from<br />

tissue samples using a Qiagen <strong>Plant</strong> Mini Kit (QIAGEN,<br />

Inc., Valencia, CA) with minor modifications in the protocol<br />

to achieve a higher concentration of DNA.<br />

AFLP analysis was carried out following Vos et al<br />

(1995) with minor modifications. The enzymes EcoRI<br />

and MseI were used for DNA digestion. Each plant<br />

sample was fingerprinted with six primer combinations.<br />

The primer extensions used were EcoAA/MseA,<br />

EcoAA/MseG, EcoAA/MseT, EcoAC/MseA, EcoAC/<br />

MseG, and EcoAC/MseT. Fragment separation and detection<br />

was carried out on a LI-COR 4300 DNA Analysis<br />

System (LI-COR Biosciences, Lincoln, NE) on a<br />

6.5% polyacrylamide gel. Only unambiguous bands<br />

(50 – 350 bp) were scored for presence or absence.<br />

Bands that were monomorphic among all samples were<br />

discarded from analysis of polymorphic bands. Principal<br />

components analysis was performed on the complete<br />

data set, on data from the three close congeners alone,<br />

on data from P. argillacea alone, and on data from the<br />

Tucker half-sib families alone. We used SAS software<br />

(SAS Institute, Cary, NC) for the analysis.<br />

RESULTS<br />

AFLP analysis produced a total of 535 reliably reproducible<br />

bands, 124 of which were polymorphic. Seventy-five<br />

of these bands were polymorphic only between<br />

P. crenulata and the other Phacelia species (P. glandulosa,<br />

P. argylensis, and P. argillacea; Table 1). This<br />

clearly demonstrated that the P. glandulosa group is<br />

strongly genetically differentiated from P. crenulata, the<br />

putative distant congener in the study. The three close<br />

congeners were much more genetically similar. Phacelia<br />

argillacea exhibited nine bands that were polymorphic<br />

with P. argylensis, seven polymorphic bands<br />

within P. glandulosa from herbarium material, and<br />

seven polymorphic bands within P. glandulosa collected<br />

by Frank Smith in western Colorado. An unexpected<br />

result was that the Smith collections were even more<br />

differentiated from other P. glandulosa than was P. argillacea,<br />

with 15 bands polymorphic between the two<br />

groups. In contrast, P. argylensis was closely similar to<br />

the herbarium-collected P. glandulosa group, with only<br />

3 polymorphic bands. Within P. argillacea, we observed<br />

a total of 30 polymorphic bands, however no polymorphic<br />

bands were found between the Tucker and<br />

Railroad populations, suggesting low genetic differentiation.<br />

When we analyzed data from all four Phacelia species<br />

included in the study, the first principal component<br />

represented 79% of the total variation and clearly separated<br />

P. crenulata form the other three species, reflect-<br />

Table 1. <strong>Number</strong> of polymorphisms identified using the AFLP (amplified fragment length polymorphism)<br />

technique between or within pairs of species or, in the case of P. glandulosa, within-species groups.<br />

P. argylensis P. glandulosa (H) P. glandulosa (F) P. crenulata<br />

P. argillacea 9 7 7 78<br />

P. argylensis 3 18 93<br />

P. glandulosa (H) 15 87<br />

P. glandulosa (F) 83<br />

(H) Herbarium-collected samples from individual herbarium specimens.<br />

(F) Field-collected bulk samples from two closely adjacent populations.<br />

129


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

ing the distant relationship between it and the other<br />

Phacelia samples. This was a consequence of the large<br />

number of AFLP bands that were polymorphic between<br />

P. crenulata and the other three species (Table 1).<br />

PCA was then used to analyze the relationships between<br />

P. argylensis, P. glandulosa, and P. argillacea.<br />

The first two principal components represented 9.2 and<br />

7.2% of the total variation. The P. argillacea samples<br />

clearly grouped separately from samples of the other<br />

two species (Figure 2). In contrast, P. argylensis<br />

grouped closely with the herbarium-collected P. glandulosa<br />

group, calling its status as a separate species into<br />

question. The field-collected western Colorado P. glandulosa<br />

was most distant from the herbarium-collected P.<br />

glandulosa group, as was also indicated by the large<br />

number of polymorphic bands between these two sets of<br />

collections, suggesting that it perhaps represents an undescribed<br />

taxon within the group (Table 1).<br />

When PCA was applied to data from the P. argillacea<br />

samples from both populations and among different<br />

years, the first two principal components, which explained<br />

7.5 and 2.9% of the total variation, provided<br />

enough separation to compare populations and years<br />

(Figure 3). The resulting data grouped each sample with<br />

cohorts from the same year more closely than by population.<br />

For example, there was no overlap between<br />

Tucker 2006 samples and the 2004 and 2008 samples<br />

from the same population. Similarly, there was no overlap<br />

between Railroad 2008 samples and the 2006 and<br />

2007 samples of the same population. In addition, individuals<br />

from the Railroad population were surrounded<br />

Figure 2. Scores on the first two axes from Principal Components Analysis of AFLP (amplified fragment length<br />

polymorphism) data for three species of Phacelia. The P. glandulosa point near the lower left hand corner of the<br />

graph represents two bulked field-collected samples from closely adjacent populations; all other points represent individual<br />

plants, or multiple individuals with identical genotypes.<br />

130


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 3. Scores on the first two axes from Principal Components Analysis of AFLP (amplified fragment length<br />

polymorphism) data for P. argillacea individuals collected from each of two populations in each of three years.<br />

by Tucker individuals on the plot, showing no clear genetic<br />

differentiation between the two populations.<br />

PCA was also used to differentiate the greenhousegrown<br />

half-sib individuals from Tucker 2004. The first<br />

two principal components, even though they explained<br />

only 3.4 and 1.7% of the total variation, generally separated<br />

the samples into their respective families (Figure<br />

4). The graphs indicate that members of half-sib families<br />

tend to resemble each other more closely than samples<br />

belonging to different half-sib families.<br />

DISCUSSION<br />

Although the AFLP marker system is not well suited<br />

for phylogenetic analysis per se, it is useful for examining<br />

the degree of genetic distinctness among closely<br />

related populations and species. In this study, PCA<br />

analysis of the AFLP bands suggests that P. argillacea<br />

is distinct from both its closest congeners (P. argylensis<br />

and P. glandulosa). Additionally, these results indic-<br />

ated that P. argylensis is more closely related to P. glandulosa<br />

than is P. argillacea, and may not be distinct<br />

from P. glandulosa. Our analysis also suggests that the<br />

differences within P. glandulosa as presently described<br />

may be greater than the differences between P. glandulosa<br />

and P. argillacea. A close examination of the population<br />

that was the source of the field-collected P. glandulosa<br />

samples from western Colorado may reveal that<br />

these samples represent a distinct and previously undescribed<br />

taxon.<br />

The AFLP analysis revealed that P. argillacea appears<br />

to have a surprising amount of genetic diversity<br />

for a species of such limited distribution. Of the total<br />

polymorphic bands encountered, 24%, or 30 bands,<br />

were polymorphic just within P. argillacea. These<br />

polymorphisms were distributed within populations, as<br />

there were no polymorphic bands between the Tucker<br />

and Railroad populations, and the PCA showed no distinct<br />

pattern by population. Instead, the PCA of P.<br />

131


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

4<br />

Family 7<br />

4<br />

Family 11<br />

4<br />

Family 14<br />

4<br />

Family 16<br />

2<br />

2<br />

2<br />

2<br />

0<br />

0<br />

0<br />

0<br />

-2<br />

-2<br />

-2<br />

-2<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

Family 17<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

Family 21<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

Family 24<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

Family 26<br />

PRINCIPAL COMPONENT 2<br />

0<br />

-2<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

0<br />

-2<br />

4<br />

2<br />

0<br />

-2<br />

Family 27<br />

-4<br />

-4<br />

-6 -4 -2 0 2 4 -6 -4 -2 0 2 4<br />

Family 39<br />

-4<br />

-6 -4 -2 0 2 4<br />

0<br />

-2<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

0<br />

-2<br />

4<br />

2<br />

0<br />

-2<br />

Family 33<br />

Family 41<br />

-4<br />

-6 -4 -2 0 2 4<br />

0<br />

-2<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

0<br />

-2<br />

Family 35<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

0<br />

-2<br />

Family 42<br />

-4<br />

-6 -4 -2 0 2 4<br />

PCA 2<br />

0<br />

-2<br />

-4<br />

-6 -4 -2 0 2 4<br />

4<br />

2<br />

0<br />

-2<br />

Family 37<br />

-4<br />

-6 -4 -2 0 2 4<br />

One Individual<br />

Two Individuals<br />

Three Individuals<br />

Four Individuals<br />

Five Individuals<br />

PRINCIPAL COMPONENT 1<br />

Figure 4. Scores on the first two axes from Principal Components Analysis of AFLP (amplified fragment length<br />

polymorphism) data for greenhouse-grown individuals belonging to 15 half-sib familes collected from the P. argillacea<br />

Tucker population in 2004. Point size reflects number of individuals with identical AFLP genotypes.<br />

argillacea indicated that individuals tended to group<br />

together by year much more than by population. These<br />

data suggest a persistent seed bank, which means a fraction<br />

of the seeds not only remain in the soil, but are viable<br />

for at least one year after production (Thompson<br />

and Grime 1979). A site characterization study of P.<br />

argillacea supported this hypothesis by suggesting that<br />

the seed bank reservoir contains an accumulation of<br />

seeds from many different years (Armstrong 1992).<br />

Preliminary results from a long term seed retrieval study<br />

with P. argillacea also support the existence of a longlived<br />

seed bank in this species. Few or no seeds have<br />

germinated in the field during the first two years, and<br />

most are still in a state of primary dormancy (Meyer<br />

unpublished data). This type of seed bank structure has<br />

also been reported in Phacelia secunda. Seeds from P.<br />

secunda were collected and allowed to germinate, and<br />

after three years a considerable fraction of the seeds remained<br />

viable but ungerminated (Cavieres 2001).<br />

PCA also suggests that collections from a single year<br />

and population of P. argillacea would depict a very narrow<br />

genetic diversity within the organism. A persistent<br />

seed bank can function as a genetic memory by accumulating<br />

seed genotypes from different years (Cabin et al.<br />

1998). In the case of the rare annual Clarkia springvillensis,<br />

analysis of seed bank samples illustrated significantly<br />

higher within-seed bank genetic diversity when<br />

compared to the adult population (McCue and Holtsford<br />

1998). The same could be true for P. argillacea, as evidenced<br />

by the pattern seen in the PCA (Figure 3). The<br />

seed bank must have a higher genetic diversity than the<br />

established plants in any one year, because of the wide<br />

range of diversity seen when comparing years. A similar<br />

situation was found in Phacelia dubia, which has small<br />

132


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

population size and was thought to suffer from genetic<br />

drift and bottlenecking. However, analysis of the seed<br />

bank and adult plants from different years showed that<br />

no alleles were lost. The seed bank was found to store<br />

the full range of different genotypes (del Castillo 1994).<br />

Our study suggests that P. argillacea exhibits this type<br />

of age-structured seed bank and genetic pattern. In addition,<br />

the age and genotype of a seed may play a part in<br />

permitting it to germinate and establish in a particular<br />

kind of year. Allowing only certain genotypes to germinate<br />

each year would produce a pattern similar to the<br />

one found in Figure 3, with little or no overlap in genotypes<br />

between years.<br />

CONCLUSIONS<br />

In the current reintroduction efforts with P. argillacea,<br />

by selecting seeds collected in just one year we<br />

may be severely limiting the genetic base of this species.<br />

As the data from this study and similar studies suggest,<br />

in cases of a persistent seed bank, the parents of<br />

each year’s crop can differ from the seedling cohorts<br />

found in the years before and after (Figure 4). By using<br />

only greenhouse-grown seeds produced from the Tucker<br />

2004 seed collection, we were inadvertently selecting<br />

for only a few specific genotypes. With individuals from<br />

just one year, the reintroduction program will almost<br />

surely suffer from inbreeding and genetic bottlenecks.<br />

To broaden the genetic base of this organism and allow<br />

for establishment of successful new populations of P.<br />

argillacea, the reintroduction program needs to include<br />

collections from several years and from both populations.<br />

ACKNOWLEDGMENTS<br />

This work was carried out with the aid of funding<br />

from the US Fish and Wildlife Service Preventing Extinction<br />

Program. We thank Heather Barnes (USFWS),<br />

Denise Van Keuren (formerly of the Uinta National Forest),<br />

Renee Van Buren and Kimball Harper (<strong>Utah</strong> Valley<br />

University), Katie Temus Merill and Ben Brulotte<br />

(Brigham Young University), Wendy Yates, Jennifer<br />

Lewinsohn, and Rita Dodge (Red Butte Gardens), Frank<br />

Smith (Western Ecological Services), and Bitsy Shultz<br />

and Susan Garvin Fitts (<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>) for<br />

logistical assistance. Permission to collect tissue samples<br />

from specimens at the Brigham Young University<br />

Herbarium is also gratefully acknowledged.<br />

LITERATURE CITED<br />

Armstrong, V.R. 1992. Site characteristics and habitat<br />

requirements of the endangered Clay Phacelia<br />

(Phacelia argillacea Atwood, Hydrophyllaceae). M.S.<br />

thesis. Brigham Young University.<br />

Atwood, N.D. 1975. A revision of the Phacelia<br />

Crenulatae group (Hydrophyllaceae) for North America.<br />

Great Basin Naturalist 35: 1-190.<br />

Cabin, R.J., R.J. Mitchell, and D.L. Marshall. 1998.<br />

Do surface plant and soil seed bank populations differ<br />

genetically? A multipopulation study of the desert mustard<br />

Lesquerella fendleri (Brassicaceae). American<br />

Journal of Botany 85(9): 1098-1109.<br />

Cavieres, L.A. and M.T.K. Arroyo. 2001. Persistent<br />

soil seed banks in Phacelia secunda (Hydrophyllaceae):<br />

experimental detection of variation along an altitude<br />

gradient in the Andes of central Chile (33°S). Journal of<br />

Ecology 89: 31-39.<br />

del Castillo, R.F. 1994. Factors influencing the genetic<br />

structure of Phacelia dubia, a species with a seedbank<br />

and large fluctuations in population size. Heredity<br />

72: 446-458.<br />

Garrison, L. 2007. Phylogenetic relationships in Phacelia<br />

(Boraginaceae) inferred from nrITS sequence data.<br />

M.S. thesis. San Francisco State University.<br />

Gilbert, C., J. Dempcy, C. Ganong, R. Patterson, and<br />

G.S. Spicer. 2005. Phylogenetic relationships within<br />

Phacelia subgenus Phacelia (Hydrophyllaceae) inferred<br />

from nuclear rDNA ITS sequence data. Systematic Botany<br />

30: 627-634.<br />

Kang, M., Q. Ye, and H. Huang. 2005. Genetic consequence<br />

of restricted habitat and population decline in<br />

endangered Isoetes sinensis (Isoetaceae). Annals of Botany<br />

96: 1265-1274.<br />

McCue, K.A. and T.P. Holtsford. 1998. Seed bank<br />

influences on genetic diversity in the rare annual Clarkia<br />

springvillensis (Onagraceae). American Journal of<br />

Botany 85: 30-36.<br />

Ronikier, M. 2002. The use of AFLP markers in conservation<br />

genetics – a case study on Pulsatilla vernalis<br />

in the Polish Lowlands. Cellular and Molecular Biology<br />

Letters 7: 677-684.<br />

Thompson, K and J.P. Grime. 1979. Seasonal variation<br />

in the seed banks of herbaceous species in ten contrasting<br />

habitats. Journal of Ecology 67: 893-921.<br />

U.S. Fish and Wildlife Service. 1978. Determination<br />

of five plants as endangered species. Federal Register<br />

43: 44809-44812.<br />

U.S. Fish and Wildlife Service. 1989. Clay phacelia,<br />

Phacelia argillacea Atwood, recovery plan. USFWS<br />

Region 6, Denver Colorado. iii + 19 pp.<br />

Vos, P. et al. 1995. AFLP: a new technique for DNA<br />

fingerprinting. Nucleic Acids Research. 23: 4407-4414.<br />

Welsh, S.L., N.D. Atwood, S. Goodrich, and L.C.<br />

Higgins. 2003. A <strong>Utah</strong> Flora, third edition. Brigham<br />

Young University, Provo, UT.<br />

133


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

A Taxonomic Revision of Astragalus lentiginosus var. maricopae<br />

and Astragalus lentiginosus var. ursinus<br />

Two Taxa Endemic to the Southwestern United States<br />

Jason A. Alexander, curator<br />

<strong>Utah</strong> Valley University Herbarium, <strong>Utah</strong> Valley University, Orem, UT<br />

and research associate, Wesley E. Niles Herbarium, University of Nevada, Las Vegas<br />

Abstract. Two taxa in the Astragalus lentiginosus complex of Section Diphysi, Astragalus lentiginosus var. maricopae<br />

and A. lentiginosus var. ursinus, have been historically overlooked by taxonomists and have had an uncertain<br />

taxonomic status. Astragalus lentiginosus var. maricopae is a highly endangered endemic (likely totaling less than<br />

5,000 individuals primarily due to habitat loss from development) and confined to a small region of igneous and granitic<br />

alluvial fans in the vicinity of Scottsdale and the Verde River drainage in northern Maricopa Co., Arizona. The<br />

second variety, A. lentiginosus var. ursinus, is a highly restricted limestone talus endemic (totaling less than 5,000<br />

individuals) and is confined to a small region of the Beaver Dam Mountains in Mohave Co., Arizona and Washington<br />

Co., <strong>Utah</strong>. Two morphological principal coordinates analyses (PCoA) were used on vouchers of these two varieties<br />

and nearly 150 specimens from related taxa in Section Diphysi. The results of the first PCoA showed that the floral<br />

and pod morphology of A. lentiginosus var. maricopae contributed highly to its distinctiveness when compared to<br />

other varieties, especially A. lentiginosus var. wilsonii (its geographically closest relative ). These results combined<br />

with field observations indicate that A. lentiginosus var. maricopae is a morphologically unique and highly endangered<br />

taxon that is threatened by disturbance and development throughout its known range. Based on the second<br />

PCoA, A. lentiginosus var. ursinus trends toward smaller pods and flowers than its geographically nearest relative (A.<br />

lentiginosus var. mokiacensis) and is herein recognized at the varietal level. Astragalus lentiginosus var. ursinus is<br />

more ecologically specialized than A. lentiginosus var. maricopae. However, most of the population is in a wilderness<br />

area and is threatened by recreational activities, not extirpation by suburban development.<br />

Marcus E. Jones (1923) was one of the first taxonomists<br />

to attempt to write a comprehensive treatment of<br />

species previously considered related to Astragalus lentiginosus<br />

Douglas ex Hook., a species described almost<br />

a century earlier by Hooker (1831). After Hooker's publication,<br />

new species were added to this complex by Asa<br />

Gray (1849, 1865) and Sereno Watson (1871), but the<br />

complex remained poorly known until the late 19th century.<br />

In 1898, Jones proposed a set of new combinations,<br />

placing some species from Section Diphysi A.<br />

Gray (sensu Gray 1863) as varieties within a greatly<br />

expanded concept of Astragalus lentiginosus (Jones<br />

1898). Jones' concept of A. lentiginosus remained relatively<br />

unchanged and culminated in his Revision of<br />

North-American Species of Astragalus (Jones 1923)<br />

which was ignored by taxonomists for two decades<br />

(Barneby 1964). Jones' core varietal concepts in Astragalus<br />

lentiginosus are largely accepted today, mainly due<br />

to the eloquence and precision of Rupert C. Barneby in<br />

his 1945 and 1964 monographs.<br />

Barneby (1945) was the first to make explicit and<br />

unambiguous the relationships between the 40 varieties<br />

of Astragalus lentiginosus and their placement into Astragalus<br />

Section Diphysi. Although many new varieties<br />

would be described and old ones further refined over the<br />

next 60 years (see Barneby 1956, 1989, Isely 1998,<br />

134<br />

Kearney & Peebles 1960, Munz & Keck 1959, Welsh<br />

1978, 1993, 2003), the recognition of 42 varieties remains<br />

even in the latest monograph (Welsh 2007). Despite<br />

the many revisions of this complex, the best description<br />

of the extremes of the diversity within the A.<br />

lentiginosus complex was published by Barneby in his<br />

first monograph:<br />

"The varieties of A. lentiginosus, as known at<br />

present, are not of equal stature: some, indeed,<br />

are doubtfully distinct, while others appear to<br />

be isolated and might, in another group of the<br />

genus, pass as species of the first rank. It is noticeable,<br />

however, that every example of the latter<br />

type is comparatively little known, whereas<br />

all those represented by extensive collections<br />

are found to intergrade at some point with a<br />

related variety" (1945: 70).<br />

The Astragalus lentiginosus complex can be divided<br />

into two major groups based on pod morphology. The<br />

first group, comprising the majority of the varieties of<br />

this complex, is distinguished by the presence of a deciduous,<br />

bladdery inflated, biloculate, ovoid to orbicular<br />

pod. These characters were viewed as the most representative<br />

by Barneby (1964) and used to distinguish


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Section Diphysi from other putatively related sections.<br />

The thickness or texture of the valve walls, the type and<br />

distribution of pubescence on the valves, and the degree<br />

of closure of the locules by the septum (if it is complete<br />

and fused to the funicular flange throughout or just<br />

within a portion of the body of the pod) is highly variable<br />

throughout the range of this species.<br />

The second group within this complex is characterized<br />

by having scarcely inflated (cylindrical to ventricose<br />

in shape and slightly inflated dorsally, if at all),<br />

thick papery to leathery, elliptic, narrowly oblong, to<br />

linear pods. The septum is generally incomplete in this<br />

group, either semi-bilocular (the septum partially divides<br />

the two locules) or sub-unilocular (the septum is<br />

less than half the width of the locule). Unlike the first<br />

group, the pods are either deciduous or long-persistent.<br />

These scarcely inflated taxa were first comprehensively<br />

described by Rydberg (1929) as Section Palantia<br />

Rydb. within the genus Tium Medik., based on the similarity<br />

of the pod morphology. The remaining members<br />

of Section Diphysi were split and included in the old<br />

world genus Cystium Steven. The degree to which these<br />

characteristics define a section or species is a major<br />

source of disagreement among all monographs of this<br />

complex (Barneby 1945, 1964, Isely 1998, Jones 1923,<br />

Rydberg 1929, Welsh 2007). The long-persistent pods<br />

in one taxon, A. lentiginosus var. mokiacensis (A. Gray)<br />

M.E. Jones, was the primary character that lead Barneby<br />

(1989) to retain it within the Section Preussiani M.E.<br />

Jones, a section distantly related to Section Diphysi.<br />

Vastly different interpretations of the significance of<br />

this character have led to often disparate views of the<br />

species boundaries and delimitations surrounding these<br />

taxa (Alexander 2005, Barneby 1964, 1989, Welsh<br />

2007). When the taxa with persistent pods are delimited<br />

as varieties, Astragalus lentiginosus becomes the only<br />

documented North American species of Astragalus to<br />

have forms with both persistent and deciduous pods.<br />

All other examples proposed by taxonomists have been<br />

split at the species or the sub-sectional level in recent<br />

monographs.<br />

In the most recent revision, Alexander (2008) considered<br />

all of the scarcely inflated varieties of A. lentiginosus<br />

to be a single evolutionary lineage and referred to<br />

them collectively as the Palantia, based on Rydberg's<br />

sectional name. The Palantia consists of A. lentiginosus<br />

var. bryantii (Barneby) J.A. Alexander, A. lentiginosus<br />

var. iodanthus (S. Watson) J.A. Alexander, A. lentiginosus<br />

var. maricopae Barneby, A. lentiginosus var.<br />

mokiacensis (including A. lentiginosus var. trumbullensis<br />

S.L. Welsh & Atwood), A. lentiginosus var. palans<br />

(M.E. Jones) M.E. Jones, A. lentiginosus var. pseudiodanthus<br />

(Barneby) J.A. Alexander, A. lentiginosus var.<br />

ursinus (A. Gray) Barneby, and A. lentiginosus var. wilsonii<br />

(Greene) Barneby. The name, Palantia, is not used<br />

herein in a nomenclatural sense as a sub-generic or sectional<br />

name. It is unusual in botany for a collective<br />

name to be required for clarity when referring to groups<br />

of morphologically similar varieties. However, with<br />

over 40 varieties and 4 or 5 lineages that are similar<br />

morphologically, a collective naming convention for A.<br />

lentiginosus is necessary to refer to these groups. As<br />

such, the Palantia is used as a convenient, informal<br />

name for the varietal group with scarcely inflated pods<br />

and is italicized following the common literary convention<br />

for unfamiliar Latin words.<br />

Within the Palantia, two taxa, Astragalus lentiginosus<br />

var. maricopae and A. lentiginosus var. ursinus have<br />

been historically overlooked by taxonomists and have<br />

an uncertain taxonomic status. Astragalus ursinus A.<br />

Gray (first reduced to a variety of A. lentiginosus by<br />

Barneby 1964) was the first to be described in 1878,<br />

along with A. mokiacensis A. Gray (Gray 1878). Gray<br />

was the first to propose a close relationship between<br />

these taxa and A. lentiginosus var. iodanthus (at that<br />

time, and until only recently, this taxon was delimited as<br />

a species). The affinity of the types of A. ursinus to extant<br />

populations has been controversial since the taxon<br />

was first described (Alexander 2005, 2008, Barneby<br />

1964, Jones 1923, Welsh 1978, 2007, Welsh & Atwood<br />

2001). In some recent taxonomic treatments (Barneby<br />

1989, Welsh 1993), the types have been regarded as an<br />

insignificant variant of A. lentiginosus var. palans.<br />

However, in others A. lentiginosus var. ursinus is recognized<br />

as an insignificant variant of A. lentiginosus var.<br />

mokiacensis (Alexander 2005, Welsh 2007, Welsh &<br />

Atwood 2001).<br />

Based on a combination of field surveys, morphometric<br />

analyses and chloroplast haplotype analyses, Astragalus<br />

lentiginosus var. ursinus was found to be genetically<br />

distinct from its geographically closest relative,<br />

A. lentiginosus var. mokiacensis (Alexander 2008, Alexander<br />

& Liston, in prep). In addition, Astragalus lentiginosus<br />

var. ursinus is a highly restricted limestone talus<br />

endemic (totaling less than 5,000 individuals), and is<br />

confined to a small region of the Beaver Dam Mountains<br />

in Mohave County, Arizona and Washington<br />

County, <strong>Utah</strong> (Alexander 2008).<br />

The second species, Astragalus lentiginosus var.<br />

maricopae, was first described in Barneby's (1945)<br />

monograph. It is often confused with A. lentiginosus<br />

var. yuccanus due to similar floral morphology (size and<br />

color) and has remained poorly known since it was first<br />

described. The floral and pod morphology are highly<br />

distinct when compared to the other members of the<br />

Palantia. If it were placed within any other section of<br />

the genus, it would be recognized at the species-level. It<br />

has universally been recognized as a variety of A. lentiginosus<br />

in all major monographs, but has only recently<br />

been found to be a restricted endemic (Alexander 2008).<br />

135


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Though A. lentiginosus var. maricopae is morphologically<br />

distinct from its geographically closest relative, A.<br />

lentiginosus var. wilsonii, it is not as genetically differentiated<br />

as A. lentiginosus var. ursinus is from A. lentiginosus<br />

var. mokiacensis (Alexander 2008, Alexander<br />

& Liston, in prep). Alexander (2008) also found that A.<br />

lentiginosus var. maricopae is a highly endangered endemic<br />

(likely totaling less than 5,000 individuals primarily<br />

due to habitat loss from suburban development)<br />

and confined to a small region of igneous and granitic<br />

alluvial fans in the vicinity of Scottsdale and the Verde<br />

River drainage in northern Maricopa County, Arizona.<br />

In this study, morphological principal coordinates<br />

analyses (PCoA), cluster analyses, and cladistic analyses<br />

are used to detect the degree of morphological differentiation<br />

between Astragalus lentiginosus var. maricopae,<br />

A. lentiginosus var. ursinus and the remaining<br />

taxa of the Palantia and whether this differentiation corresponds<br />

to species or varietal delimitations in preparation<br />

for monographic revision of the A. lentiginosus<br />

complex.<br />

MATERIALS AND METHODS<br />

Field observations and voucher specimens were<br />

made from spring 2001 to summer 2004 throughout the<br />

range of the Palantia. Most populations were visited<br />

several times and were observed during early flower,<br />

maturity, and senescence. Vouchers for this study are<br />

deposited at NY, OSC, RSA, UNLV, and UVSC.<br />

Herbarium specimens were examined at UC in <strong>December</strong><br />

of 1999, BRY in August of 2000, GH in August<br />

of 2002, NY in October of 2003, and UNLV in July of<br />

2002 and 2003. Additional herbarium specimens were<br />

obtained on loan from BRY, CAS, DS, K, POM, RM,<br />

and RSA.<br />

Specimens from taxa in Astragalus Section Diphysi<br />

were examined for two morphological PCoA studies.<br />

The first, the 153 specimen Palantia PCoA, focused on<br />

heavily sampling all members of the group and was designed<br />

to evaluate the morphological distinctiveness of<br />

A. lentiginosus var. maricopae. 103 specimens of A.<br />

lentiginosus var. palans were used representing all major<br />

regions of its range, including the type population.<br />

Also included were multiple specimens of A. lentiginosus<br />

var. bryantii (10), A. lentiginosus var. ursinus (10),<br />

and A. lentiginosus var. wilsonii (15). Due to the poor<br />

nature of most herbarium specimens of A. lentiginosus<br />

var. maricopae, 5 specimens were collected in the field<br />

in 2005 for the morphological analysis. To ensure that<br />

the range of regional variation was present in the PCoA,<br />

representative specimens (one except where noted) were<br />

included from these morphologically similar and geographically<br />

proximal varieties: A. lentiginosus var. ambiguus<br />

Barneby (the type specimen); A. lentiginosus var.<br />

araneosus (Sheld.) Barneby; a population of a A. lent-<br />

iginosus (from Chloride, Mohave County, Arizona, interpreted<br />

herein as an intermediate to A. lentiginosus<br />

var. yuccanus M.E. Jones) considered part of A. lentiginosus<br />

var. ambiguus in Barneby (1964); A. lentiginosus<br />

var. iodanthus; A. lentiginosus var. mokiacensis (2 type<br />

specimens); A. lentiginosus var. pseudiodanthus; A. lentiginosus<br />

var. stramineus (Rydb.) Barneby (the type<br />

specimen); A. lentiginosus var. vitreus Barneby (the<br />

type specimen); and A. lentiginosus var. yuccanus (the<br />

type specimen).<br />

The second, the 43 specimen Astragalus lentiginosus<br />

var. mokiacensis PCoA, focused on further evaluating<br />

the morphological distinctiveness of A. lentiginosus var.<br />

ursinus. Fifteen specimens from throughout the range<br />

of A. lentiginosus var. ursinus were examined, including<br />

the type specimens. For comparison, 28 specimens from<br />

throughout the range of A. lentiginosus var. mokiacensis<br />

(including the type specimens) were examined. Multiple<br />

specimens and types of the synonym, A. lentiginosus<br />

var. trumbullensis S.L. Welsh & N.D. Atwood were<br />

included in this study. These data are a modified version<br />

of the data used in Alexander (2005).<br />

In both PCoA's, duplicate specimens or specimens<br />

from the same locality were used to determine the character<br />

states of missing data. Qualitative characters that<br />

were found to be polymorphic within a single individual<br />

were excluded.<br />

The morphological matrices for both studies were<br />

transformed into a Gower (1971) similarity matrix, a<br />

process that is not sensitive to data sets with mixed ordinal,<br />

nominal, continuous, and ratio data types. The matrix<br />

was then used in the PCoA. A Kendall's tau correlation<br />

between the PCoA axes and all morphological characters<br />

was used to determine the magnitude of the contribution<br />

of characters to the overall analysis (Easdale et<br />

al. 2007, Hammer et al. 2001). All correlations and<br />

PCoA analyses were performed using Paleontological<br />

Statistics (PAST) version 1.76 (Hammer et al. 2001). A<br />

Euclidean distance dendrogram was also created using<br />

PAST for the cluster analysis.<br />

PAUP* for Windows version 4.0 beta 10 (Swofford,<br />

2002) was used to assess the relationships among 30<br />

specimens of the Palantia and Section Diphysi using<br />

cladistic methodologies.<br />

The PCoA, cluster, and parsimony analyses were<br />

used to address the following questions:<br />

1) Do populations of A. lentiginosus var. maricopae and<br />

A. lentiginosus var. ursinus form groups discrete from<br />

the other members of the Palantia ?<br />

2) Which morphological characters contribute to the<br />

observed groups?<br />

3) Are these groups morphologically differentiated from<br />

the closely related and geographically proximal varieties<br />

of A. lentiginosus?<br />

136


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

RESULTS<br />

Distribution maps (Figures 1, 2) show the localities<br />

of specimens of Astragalus lentiginosus var. maricopae<br />

and A. lentiginosus var. ursinus examined for this study.<br />

Specific vouchers labeled in figures are shown in Table<br />

1. A detailed list of the vouchers examined can be found<br />

in the Taxonomic Treatment and in Appendix 1.<br />

Of the 24 morphological characters examined for the<br />

153 specimen Palantia PCoA, two were constant and<br />

not used (Table 2), one was discarded due to character<br />

state scoring issues, and 22 were variable. The first<br />

component of the PCoA explained 22.1% of the total<br />

variance (Table 3). The largest correlations to the first<br />

axis were from pod raceme orientation (podro), degree<br />

of pod incurve (podpi), pod pedicel orientation (podpo),<br />

banner color (bannc), and pod persistence (poddp). The<br />

second component of the PCoA explained 11.8% of the<br />

total variance. The largest correlations to the second<br />

axis were from pod pubescence (podpu), leaf abaxial<br />

pubescence (leafab), pod shape in cross section (podsc),<br />

and pod inflation (podin). All other components of the<br />

PCoA explained less than 10% of the total variance. The<br />

scatterplot of the first two components of the PCoA is<br />

shown in Figure 3.<br />

Of the 24 morphological characters examined for the<br />

43 specimen Astragalus lentiginosus var. mokiacensis<br />

PCoA, 12 were found to be variable in this subset of the<br />

data (Table 4). The first component of the PCoA explained<br />

48.4% of the total variation (Table 5). The largest<br />

correlations to the first axis were from pod shape in<br />

Figure 1. Distribution map of Astragalus lentiginosus<br />

var. maricopae in Arizona. The distributions of Astragalus<br />

lentiginosus var. bryantii and A. lentiginosus var.<br />

wilsonii are shown for reference. Table one contains the<br />

legend of letter codes used for specific vouchers shown.<br />

Vouchers and major populations are labeled as follows:<br />

(caret), A. lentiginosus var. bryantii; (upward triangle),<br />

A. lentiginosus var. maricopae; (circle), A. lentiginosus<br />

var. wilsonii; (X), Cameron population of A. lentiginosus<br />

var. wilsonii putatively intermediate to A. lentiginosus<br />

var. palans.<br />

Figure 2. Distribution map of Astragalus<br />

lentiginosus var. ursinus. The distribution<br />

of A. lentiginosus var. mokiacensis<br />

is shown for reference. Table 1 contains<br />

the legend of letter codes used for<br />

specific vouchers shown. Taxa are labeled<br />

as follows: (X) A. lentiginosus var.<br />

ursinus; (upward triangle) A. lentiginosus<br />

var. mokiacensis, mokiacensis minor<br />

variant; (diamond) A. lentiginosus var.<br />

mokiacensis, Gold Butte minor variant;<br />

(caret) A. lentiginosus var. mokiacensis,<br />

trumbullensis minor variant. For a taxonomic<br />

treatment for the morphological<br />

variants shown herein for A. lentiginosus<br />

var. mokiacensis, see Alexander<br />

(2005, 2008).<br />

137


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. Specific vouchers identified in the PCoA, cluster analysis, and parsimony figures.<br />

The Map Code column identifies vouchers on the distribution maps and PCoA figures. Label data for vouchers of A.<br />

lentiginosus var. maricopae and A. lentiginosus var. ursinus can be found in the Taxonomic Treatment and the data<br />

for all other varieties can be found in Appendix 1. Vouchers collected by Alexander with letters in brackets (i.e. Alexander<br />

2367 [A]) are specific individuals on vouchers measured for the morphological analyses. Type: H=Holotype,<br />

L=Lectotype, I=Isotype, IL=Isolectotype, P=Paratype, v=type vicinity<br />

Taxon Type Map<br />

Code<br />

Herbarium<br />

Voucher<br />

A. lentiginosus var. ambiguus H, I B RSA, CAS Ripley & Barneby 3403<br />

A. lentiginosus var. ambiguus<br />

(intermediate to var. yuccanus)<br />

D OSC, UNLV Alexander 2325<br />

A. lentiginosus var. araneosus I A NY, ORE, GH Jones s.n. from June 1880;<br />

Jones 1807 from June 1880<br />

A. lentiginosus var. bryantii H N CAS Bryant s.n. 15 Dec. 1939<br />

A. lentiginosus var. bryantii N 2 GH Holmgren et al. 15609<br />

A. lentiginosus var. iodanthus v J NY, ORE Jones s.n. from May 1882;<br />

Jones 3837 from May 1882<br />

A. lentiginosus var. maricopae v S OSC, UNLV Alexander 1621 [A]<br />

A. lentiginosus var. maricopae v S 1 OSC, UNLV Alexander 1621 [C]<br />

A. lentiginosus var. maricopae H S 2 US Harrison 1790<br />

A. lentiginosus var. mokiacensis<br />

(putative type locality)<br />

A. lentiginosus var. mokiacensis<br />

(published type locality)<br />

L, IL L GH, NY Palmer 105<br />

L,IL L 2 GH, NY Palmer 105<br />

A. lentiginosus var. mokiacensis Q NY, POM Jones 5058<br />

A. lentiginosus var. palans H,I E POM, GH Eastwood s.n. June 1892<br />

A. lentiginosus var. palans<br />

(type of A. amplexus)<br />

H,I F RM, GH Payson 335<br />

A. lentiginosus var. palans P G GH Eastwood s.n. May 1892<br />

A. lentiginosus var. palans H NY Barneby 13104<br />

A. lentiginosus var. palans H 1 NY Demaree 43807b<br />

A. lentiginosus var. palans H 2 NY Holmgren & Holmgren 12796<br />

A. lentiginosus var. palans H 3 NY, POM,<br />

BRY<br />

Jones 5218<br />

A. lentiginosus var. palans H 4 NY Jones 5218a<br />

A. lentiginosus var. palans H 5 NY Raven 13079<br />

A. lentiginosus var. palans H 6 NY Ripley & Barneby 8662<br />

A. lentiginosus var. palans H 7 NY Weber 4735<br />

A. lentiginosus var. pseudiodanthus v K OSC, UNLV Alexander 1631<br />

A. lentiginosus var. stramineus H,I T NY, GH Palmer s.n. in 1870<br />

138


A. lentiginosus var. ursinus<br />

(published type locality)<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1, continued<br />

Taxon Type Map<br />

Code<br />

Herbarium<br />

Voucher<br />

L,IL M GH, NY Palmer s.n. 1877<br />

A. lentiginosus var. ursinus M 1 OSC, UNLV Alexander 2120 [A]<br />

A. lentiginosus var. ursinus<br />

(Mokiak Pass elements mounted with the<br />

type by Gray)<br />

M 2 GH Palmer s.n. 1877<br />

A. lentiginosus var. vitreus H,I U POM, NY Maguire & Blood 4413<br />

A. lentiginosus var. wilsonii v R OSC, UNLV Alexander 2367 [A]<br />

A. lentiginosus var. wilsonii v R 2 OSC, UNLV Alexander 2334 [D]<br />

A. lentiginosus var. wilsonii (type locality) H,I R 3 ND Wilson s.n. from May 1893<br />

A. lentiginosus var. wilsonii<br />

(population intermediate with var. bryantii,<br />

var. mokiacensis, or var. ursinus)<br />

A. lentiginosus var. wilsonii<br />

(population putatively intermediate to var.<br />

palans)<br />

P CAS Eastwood 5748<br />

P 2 NY Demaree 43807<br />

A. lentiginosus var. yuccanus H C POM Jones 3886<br />

cross section (podsc), pod shape in longitudinal section<br />

(podsl), pod pedicel orientation (podpo), leaf abaxial<br />

pubescence (leafab), keel length (keell), calyx teeth<br />

shape (calyxs), and pod orientation on raceme. The second<br />

axis of the PCoA explained 11.7% of the total variation.<br />

The largest correlation to the second axis was<br />

from leaf adaxial pubescence (leafad). Moderate correlations<br />

were obtained for pod stipe length (pods), calyx<br />

teeth shape (calyxs), and wing color (wingc). A scatterplot<br />

of the first two coordinates of the PCoA is shown in<br />

Figure 4.<br />

In PAUP*, an heuristic search of 100 random addition<br />

sequences with TBR branch swapping was started<br />

with the data set of 21 morphological characters from 30<br />

specimens of the Palantia and related members of A.<br />

lentiginosus. All 21 characters were parsimony informative.<br />

Sixteen most parsimonious trees of length 110<br />

were recovered (HI = 0.5636; RI = 0.6416; CI = 0.4364;<br />

RC= 0.2800). Figures 5-8 are the strict consensus of<br />

trees of length 110 with major characters state changes<br />

mapped on the clades. The clades in this analysis have<br />

low support. A bootstrap analysis of 10,000 replicates<br />

resulted in only five clades having 70% or higher support<br />

(pseudiodanthus & iodanthus, clade A, 73%;<br />

vitreus 4413 to yuccanus 3886, clade B, 73%; yuccanus<br />

3886 & ambiguus 2325, clade C, 87%; wilsonii 2367A<br />

& 2334D clade, 75%; maricopae 1621A & 1621C,<br />

clade E, 95%). Only banner color (Figure 5), pod persistence<br />

(excluding the reversal to a deciduous pod in A.<br />

lentiginosus var. wilsonii, Eastwood 5748; Figure 6),<br />

pod raceme orientation (Figure 7) and degree of pod<br />

incurve (not shown) had a high consistency with little or<br />

no character state reversals.<br />

A dendrogram (Figure 9) of a Euclidean similarity<br />

matrix obtained from a cluster analysis showed nearly<br />

the same topology as the tree obtained from the parsimony<br />

analysis.<br />

DISCUSSION<br />

Outgroup selection in this study proved to be problematic.<br />

Barneby (1964) proposed that a plant similar to<br />

the small flowered Astragalus lentiginosus var. salinus<br />

(a taxon with bladdery inflated pods) was the ancestor to<br />

the members of the A. lentiginosus complex and that<br />

this complex was closely related to Section Inflati A.<br />

Gray, a large species complex with unilocular, bladdery<br />

inflated pods. Nuclear inter-transcribed spacer (ITS)<br />

DNA sequence data have shown that A. lentiginosus has<br />

an identical sequence to that of A. purshii Douglas ex<br />

Hook. (and an almost identical sequence to that of A.<br />

utahensis (Torr.) Torr. & A. Gray) of Section Argophylli<br />

A. Gray (a section composed primarily of taxa with<br />

139


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 2. List of 24 morphological characters and<br />

their abbreviations used in the 153 specimen<br />

Palantia PCoA. Two were constant (C: calyxd, calyxo)<br />

and not used in the PCoA, parsimony or cluster<br />

analysis. One (O: calyxs) was not used due to the finding<br />

that it was variable on the same plant (and within<br />

most calyces). For a list of character states for the characters<br />

below, see Appendix 2. Characters were coded as<br />

multistate continuous variation (R), binary state (B), or<br />

multistate (M).<br />

1. Adaxial leaflet pubescence (Leafad) M<br />

2. Abaxial leaflet pubescence (Leafab) M<br />

3. Leaf and stem hair length (leafh) R<br />

4. Leaflet number (leafn) M<br />

5. Inflorescence length in flower (inflw) R<br />

6. Calyx tube length (calyxl) R<br />

7. Calyx pubescence density (calyxd) M, C<br />

8. Calyx teeth shape (calyxs) M, O<br />

9. Calyx teeth orientation (calyxo) M, C<br />

10. Keel length (keell) R<br />

11. Keel color (keelc) M<br />

12. Banner color (bannc) M<br />

13. Inflorescence length in fruit (infr) R<br />

14. Pod pedicel orientation (podpo) M<br />

15. Pod length X width ratio (podr) R<br />

16. Pod deciduous or persistent (poddp) B<br />

17. Pod shape, longitudinal section (podsl) M<br />

18. Pod shape, cross section (podsc) M<br />

19. Pod orientation on raceme (podro) M<br />

20. Pod orientation, degree of pod incurve (podpi) R<br />

21. Pod inflation (podin) M<br />

22. Pod valve texture (podt) M<br />

23. Pod pubescence (podpu) M<br />

24. Pod valve color (podvc) M<br />

scarcely inflated, unilocular, leathery, deciduous pods),<br />

and not to members of Section Inflati (4-6 base pair divergence;<br />

Alexander, unpublished data, Wojciechowski<br />

et al. 1993, 1999). The ITS sequence data suggest that a<br />

deciduous, unilocular, leathery, scarcely inflated pod is<br />

the putative ancestral state in this complex. Based, in<br />

part, on these data, members of Section Argophylli were<br />

used as outgroups for chloroplast haplotype analyses in<br />

Knaus (2008). Taxa in other putatively closely related<br />

sections (Section Monoenses Barneby, Section Cystiella<br />

Barneby, Section Circumdati (M.E. Jones) Barneby, or<br />

Section Platytropides Barneby; all of which have taxa<br />

with inflated pods) have not been fully investigated in<br />

molecular analyses. The selection of any member of<br />

Section Argophylli as an outgroup automatically polarizes<br />

the ancestral state of the group as a unilocular or<br />

partially bilocular, scarcely inflated, deciduous pod. In<br />

addition, haplotypes within the Argophylli sampled by<br />

Knaus (2008) were found to be nearly twenty steps<br />

more distant from the haplotypes examined in the<br />

Palantia. Finding that A. lentiginosus var. iodanthus and<br />

A. lentiginosus var. pseudiodanthus (both of which have<br />

been universally delimited as species until Alexander<br />

2009), are not highly genetically or morphologically<br />

distinct from A. lentiginosus is a problem for outgroup<br />

selection in this study, primarily with the parsimony<br />

analysis and the genetic analyses (Alexander & Liston,<br />

in prep). The terminal taxa in this study are also not recognized<br />

at the species level, which violates assumptions<br />

in parsimony analyses. As a result, robust phylogenetic<br />

conclusions cannot be made with these data.<br />

Instead, the cladistic analyses were used to investigate<br />

patterns of character state changes within the<br />

Palantia. Astragalus lentiginosus var. iodanthus, and A.<br />

lentiginosus var. pseudiodanthus were selected as outgroups<br />

based on their relatively greater genetic and morphologic<br />

distance from the Palantia based on the results<br />

from Alexander (2008), Knaus (2008), and Alexander &<br />

Liston (in prep). Even if the inflated varieties related to<br />

A. lentiginosus var. yuccanus were used as outgroups<br />

for the Palantia, the morphological trends discussed<br />

herein would not change (see Figures 5-8) since the major<br />

clades of the Palantia would still be split in two<br />

clades. More thorough molecular analyses of species<br />

potentially closely related to Section Diphysi (with both<br />

inflated and scarcely inflated pods) are needed before a<br />

robust molecular and morphological phylogenetic study,<br />

with a satisfactory outgroup, can be attempted.<br />

Despite this, conclusions of the taxonomic status of<br />

Astragalus lentiginosus var. maricopae and A. lentiginosus<br />

var. ursinus can be made. The results of the PCoA<br />

analyses and the genetic analyses (Alexander & Liston,<br />

in prep) indicate that species level delimitations for<br />

many of the Palantia have much weaker support than<br />

previously thought by Alexander (2005). Astragalus<br />

140


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 3. Results of the 153 specimen Palantia PCoA. This analysis used 21 variable characters. Each is<br />

listed in the Kendall rank correlations below.<br />

Eigenvalues<br />

1 2<br />

1.36 0.72<br />

Percent of Total Variance Explained<br />

22.06 11.75<br />

Kendall rank correlations and probabilities<br />

(between PCoA coordinates and morphological characters, significance of p


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 3. Scatterplot of the 153 specimens Palantia PCoA using 21 variable, morphological characters. The first and<br />

second axes represent 33.5% of the variation. Taxa are labeled as: (circle), Astragalus lentiginosus var. palans including<br />

the Cameron, Arizona specimens; (caret), A. lentiginosus var. bryantii; (upward triangle), A. lentiginosus var.<br />

maricopae; (X), A. lentiginosus var. ursinus; (+), A. lentiginosus var. wilsonii including South Rim, Arizona specimens.<br />

Relevant specimens of these and other taxa (diamonds) are labeled with letters. Table one contains the legend<br />

of letter codes used for specific vouchers shown.<br />

Table 4. List of 12 morphological characters from<br />

the 43 specimen Astragalus lentiginosus var.<br />

mokiacensis PCoA. This analysis used a modified version<br />

of the data used in Alexander (2005). For a list of<br />

character states for character below, see Appendix 2.<br />

Characters were coded as multistate continuous variation<br />

(R), binary state (B), or multistate (M).<br />

1. Adaxial leaflet pubescence (leafad) M<br />

2. Abaxial leaflet pubescence (leafab) M<br />

3. Calyx tube length (calyxl) R<br />

4. Calyx teeth shape (calyxs) M<br />

5. Keel length (keell) R<br />

6. Wing color (wingc) M<br />

7. Pod length X width ratio (podr) R<br />

8. Pod pedicel orientation (podpo) M<br />

9. Pod shape, longitudinal section (podsl) M<br />

10. Pod shape, cross section (podsc) M<br />

11. Pod orientation on raceme (podro) M<br />

12. Pod stipe length (pods) R<br />

lentiginosus var. maricopae, A. lentiginosus var.<br />

mokiacensis, and A. lentiginosus var. ursinus, all with<br />

persistent pods (an otherwise dependable species-level<br />

character in Astragalus according to Barneby 1964),<br />

were confirmed to be closely related to varieties of A.<br />

lentiginosus with deciduous pods (A. lentiginosus var.<br />

palans and A. lentiginosus var. wilsonii). A haplotype<br />

network derived from an analysis of chloroplast microsatellites<br />

(Alexander & Liston, in prep) shows A. lentiginosus<br />

var. maricopae is neither highly genetically<br />

differentiated from A. lentiginosus var. palans, A. lentiginosus<br />

var. wilsonii, nor A. lentiginosus var. ursinus.<br />

Astragalus lentiginosus var. ursinus was found to be<br />

more genetically similar to the long distance disjunct, A.<br />

lentiginosus var. wilsonii, than to its geographically<br />

proximal relative, A. lentiginosus var. mokiacensis<br />

(Alexander 2008, Alexander & Liston, in prep).<br />

Though Astragalus lentiginosus var. maricopae is<br />

not highly genetically differentiated from its geographically<br />

nearest relative, A. lentiginosus var. wilsonii, it is<br />

distinct morphologically. The PCoA analysis shows that<br />

the specimens of A. lentiginosus var. maricopae form a<br />

morphologically distinct group away from A. lentiginosus<br />

var. mokiacensis, A. lentiginosus var. wilsonii, and<br />

A. lentiginosus var. ursinus. The distance is not farther<br />

than A. lentiginosus var. pseudiodanthus is from the lar-<br />

142


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

ger cluster of A. lentiginosus var. palans. Also, the inflated<br />

members of A. lentiginosus sampled (see Figure<br />

3: A. lentiginosus var. araneosus, A; versus A. lentiginosus<br />

var. stramineus, T) in this study are also spread an<br />

equivalent distance apart. The presence of a yellow<br />

flower and cylindrical pods contributed highly to the A.<br />

lentiginosus var. maricopae group. Though the flower<br />

color of A. lentiginosus var. maricopae was reported by<br />

Barneby (1964) to be ochroleucous in the type description,<br />

field observations in 2005 and 2006 revealed that<br />

the flower is yellow to light yellow in color, but not as<br />

deep a yellow as that found in European Astragalus,<br />

Thermopsis, or Trifolium. Ochroleucous flowers in Astragalus<br />

tend to have a cream tint and dry a whitish-tan,<br />

or tend to be distinctly white, basally, and grade to a<br />

yellowish tint, apically, especially in age. The flower<br />

color, the distinctiveness of the pod morphology, and<br />

the range disjunction could be utilized as support for a<br />

species-level delimitation for this taxon. However, A.<br />

lentiginosus var. maricopae is not the only variety in<br />

this complex with yellowish flowers. Though some individuals<br />

of the southern California endemic, A. lentiginosus<br />

var. nigricalycis M.E. Jones, seem to have creamish<br />

to greenish-white flowers, most have yellow flowers<br />

that dry to a darker yellow in age. Also, A. lentiginosus<br />

var. bryantii has pods that are narrower, longer, and<br />

more tubular than those in A. lentiginosus var. maricopae<br />

(see the taxonomic treatment below for more specific<br />

morphological differences). When considering<br />

both the genetic and morphological data, A. lentiginosus<br />

var. maricopae is just one of several taxa at the extreme<br />

edge of the range of variation in A. lentiginosus and one<br />

of the most morphologically distinct varieties in the<br />

Palantia.<br />

In contrast, Astragalus lentiginosus var. ursinus is<br />

genetically distinct from its nearest relative, A. lentiginosus<br />

var. mokiacensis (Alexander 2008, Alexander &<br />

Liston, in prep). The two are, however, much more sim-<br />

Table 5. Results of 43 specimen Astragalus lentiginosus var. mokiacensis PCoA. This analysis used 12<br />

variable characters. Each is listed in the Kendall rank correlations below.<br />

Eigenvalues<br />

1 2<br />

1.78 0.43<br />

Percent of Total Variance Explained<br />

48.4 11.69<br />

Kendall rank correlations and probabilities<br />

(between PCoA coordinates and morphological characters, significance of p


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 4. Scatterplot of the 43 specimen, A. lentiginosus var. mokiacensis PCoA using 12 variable, morphological<br />

characters. The first and second axes represent 60.1 % of the variation. Taxa and variants are labeled as: (circle)<br />

mokiacensis minor variant; (X) trumbullensis minor variant; (diamond) Gold Butte minor variant; (+) A. lentiginosus<br />

var. ursinus (See Table 1 for the legend of letter codes used for specific vouchers shown). For a taxonomic treatment<br />

of the morphological variants shown herein for A. lentiginosus var. mokiacensis, see Alexander (2005, 2008).<br />

Figure 5. Strict consensus of 16 most<br />

parsimonious trees of length 110<br />

showing banner color (bannc) character<br />

state changes. The letters above<br />

branches are the only highly supported<br />

clades in the tree (above 70% bootstrap<br />

support; see the results section<br />

for values for each of these lettered<br />

clades). Table 1 contains the legend of<br />

letter codes used for specific vouchers<br />

shown to the right of the taxon labels.<br />

144


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 6. Strict consensus of 16 most<br />

parsimonious trees of length 110<br />

showing pod deciduous or persistent<br />

(podpd) character state changes. The<br />

letters above branches are the only<br />

highly supported clades in the tree<br />

(above 70% bootstrap support; see the<br />

results section for values for each of<br />

these lettered clades). Table one contains<br />

the legend of letter codes used for<br />

specific vouchers shown to the right of<br />

the taxon labels.<br />

Figure 7. Strict consensus of 16 most<br />

parsimonious trees of length 110<br />

showing pod raceme orientation<br />

(podro) character state changes. The<br />

letters above branches are the only<br />

highly supported clades in the tree<br />

(above 70% bootstrap support; see<br />

the results section for values for each<br />

of these lettered clades). Table one<br />

contains the legend of letter codes<br />

used for specific vouchers shown to<br />

the right of the taxon labels.<br />

145


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 8. Strict consensus of 16<br />

most parsimonious trees of length<br />

110 showing pod inflation (podin)<br />

character state changes. The letters<br />

above branches are the only<br />

highly supported clades in the tree<br />

(above 70% bootstrap support; see<br />

the results section for values for<br />

each of these lettered clades). Table<br />

one contains the legend of letter<br />

codes used for specific vouchers<br />

shown to the right of the taxon<br />

labels.<br />

Figure 9. Cluster analysis<br />

dendrogram of the morphological<br />

data from 30 specimens<br />

of the Palantia based on<br />

a Euclidean distance matrix<br />

calculated by PAST ver. 1.76.<br />

Table one contains the legend<br />

of letter codes used for specific<br />

vouchers shown to the<br />

right of the taxon labels.<br />

146


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

ilar morphologically than A. lentiginosus var. maricopae<br />

is to A. lentiginosus var. wilsonii. The morphological<br />

similarity and geographic proximity of A. lentiginosus<br />

var. mokiacensis and A. lentiginosus var. ursinus (the<br />

nearest populations are at least 20 km apart in adjacent<br />

mountain ranges) has been the primary evidence for<br />

placing them into the same taxon in the latest monograph<br />

(Welsh 2007). The morphological distinctiveness<br />

of A. lentiginosus var. ursinus is subtle, but it is, nevertheless,<br />

present. In the A. lentiginosus var. mokiacensis<br />

PCoA analysis, specimens of A. lentiginosus var. ursinus<br />

appear to grade into the specimens of A. lentiginosus<br />

var. mokiacensis (including the type specimens of A.<br />

lentiginosus var. trumbullensis). The parsimony and<br />

cluster analyses also show A. lentiginosus var. ursinus<br />

in the same region of the tree as A. lentiginosus var.<br />

mokiacensis. However, A. lentiginosus var. ursinus does<br />

not share haplotypes with A. lentiginosus var. mokiacensis.<br />

It shares the most haplotypes with A. lentiginosus<br />

var. wilsonii (Alexander & Liston, in prep). Based on<br />

the lack of shared haplotypes with A. lentiginosus var.<br />

mokiacensis and a trend toward smaller pods and flowers<br />

(see the taxonomic treatment below for more specific<br />

morphological differences), A. lentiginosus var.<br />

ursinus is recognized herein at the varietal level following<br />

the delimitation proposed by Barneby (1945, 1964).<br />

Taxonomic Treatment<br />

The taxonomic revision herein is a first step in a full<br />

monograph of the Astragalus lentiginosus complex. As<br />

such, the specimens used in the morphological analysis<br />

are labeled in a separate voucher list. Where applicable,<br />

a list of specimens examined by the author but not yet<br />

included in morphological analyses is also included.<br />

Species delimitations in the taxonomic revision follow a<br />

phenetic species concept (Sokal, 1973; Luckow 1995).<br />

The original goal of this study was to apply a phylogenetic<br />

species concept, however, the genetic and morphologic<br />

data obtained could not be analyzed robustly<br />

using cladistic methodologies. Primarily, population<br />

level data are largely ignored since the smallest taxonomic<br />

units of phylogenetic analyses are species (Nixon<br />

& Wheeler 1990, Cracraft 1983, Luckow 1995). Table 1<br />

is a list of notable specimens identified in maps in this<br />

revision. Following the key are complete taxonomic<br />

treatments for A. lentiginosus var. maricopae and A.<br />

lentiginosus var. ursinus. Treatments for the other taxa<br />

in the key can be found in Alexander (2008).<br />

Key to the Palantia and related varieties of Astragalus lentiginosus<br />

1. Pod, in longitudinal section, linear, lanceolate, oblong, or elliptic, the shape cylindrical and not inflated or ventricose<br />

and scarcely inflated dorsally or laterally, the valves stiffly papery to coriaceous, bilocular, semibilocular,<br />

or sub-unilocular, the septum to 2.5 mm wide and not fused to the funicular flange.<br />

2. Pods long-persistent, sessile on a minute boss on the receptacle or contracted at the base into an incipient stipe<br />

0.4 to 0.7 (1.0) mm long.<br />

3. Banner light yellow, without a central white or striped spot (immaculate); keel slightly darker than the banner,<br />

drying yellowish brown and immaculate; wings slightly darker than the banner, drying yellowish brown.<br />

…………………………………………………………………………………..A. lentiginosus var. maricopae<br />

3. Banner light to dark purple with a white & purple striate central spot; keel light purple & dark purple maculate;<br />

wings light to dark purple, sometimes with white tips.<br />

4. Pods straight or slightly incurved, 20-28 (-32) mm long, 4-7.1x longer than wide, the pedicel ascending or<br />

spreading, straight or curved; leaflets glabrous to moderately pubescent adaxially, at least sparsely<br />

pubescent abaxially; [=A. mokiacensis, A. lentiginosus var. trumbullensis]………………………………<br />

………………………………………………………………………………….A. lentiginosus var. mokiacensis<br />

4. Pods incurved, 13-20 (-23) mm long, 2.4-4.7x longer than wide; the pedicel ascending or erect, straight<br />

or curved; leaflets glabrous adaxially, glabrous or sparsely pubescent abaxially; [=A. ursinus] ……….<br />

………………………………………………………………………………………A. lentiginosus var. ursinus<br />

2. Pods deciduous (sometimes tardily so) by a cellular abscission layer between the receptacle and gynoecium,<br />

sessile on a minute boss on the receptacle or contracted at the base into an incipient stipe or gynophore 0.3 to<br />

0.5 (0.7) mm long.<br />

5. Pods semi-bilocular to nearly bilocular, the septum 1-2.5 mm wide, the body incurved 120-180°, incurved<br />

less than 90°, or straight, the valves in cross section cordate, obcordate, or terete, in longitudinal section<br />

elliptic, linear, or oblong.<br />

6. Pods falcate to hamate (nearly circular), occasionally lunate, incurved to 120-180°, the pedicel deflexed<br />

or declined…………………………………………………………………. A. lentiginosus var. palans<br />

6. Pods lunate to falcate, incurved less than 90° to nearly straight, the pedicel erect, ascending, or spreading.<br />

147


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

7. Pods spreading, 6-8.8 times longer than wide, incurved less than 90°, the valves stiff papery, in longitudinal<br />

section linear or narrowly oblong…………………………………A. lentiginosus var. bryantii<br />

7. Pods ascending to erect, 4-6 times longer than wide, incurved less than 90° or nearly straight, the<br />

valves leathery to thick leathery (subligneous), often with prominent reticulate veins, in longitudinal<br />

section narrowly elliptic or oblong ……………………………………….A. lentiginosus var. wilsonii<br />

5. Pods semi-bilocular to sub-unilocular, the septum to 1.5 mm wide, the body incurved to 180° (in most individuals<br />

nearly circular), the valves in cross section oblong, obcordate, or triangular, in longitudinal section<br />

narrowly elliptic, lanceolate, or linear.<br />

8. Stems arising from a superficial root-crown; herbage glabrous to strigulose, rarely villosulous with hairs<br />

to 1.0 mm long; habitat various; [=A. iodanthus] ………………………A. lentiginosus var. iodanthus<br />

8. Stems arising from a subterranean root-crown; herbage densely villous to villosulous with hairs 0.7-1.2<br />

mm long; habitat sandy pockets of alluvial fans and stabilized dunes; [=A. pseudiodanthus] ………..<br />

………………………………..………………………………….. A. lentiginosus var. pseudiodanthus<br />

1. Pod, in longitudinal section, ovoid to globose, the shape terete to didymous, bladdery inflated, the inflation dorsiventrally<br />

and laterally, or bladdery-ventricose and the inflation greater dorsally and laterally than ventrally,<br />

the valves papery membranous to stiffly papery (occasionally coriaceous), bilocular, the septum over 2 mm<br />

wide and weakly fused to the<br />

1. Astragalus lentiginosus var. maricopae<br />

Astragalus lentiginosus var. maricopae Barneby, Leafl.<br />

W. Bot. 4:140. 1945.<br />

TYPE: U.S.A. ARIZONA: MARICOPA CO.: roadside<br />

near Tempe, 4 May 1926, G.J. Harrison 1790<br />

(HOLOTYPE: US!). Map: Figure 1.<br />

Short lived perennial herbs, 3-8 dm tall; stems ascending,<br />

single or several in clumps from a superficial root<br />

crown; herbage glabrous to sparsely strigulose with basifixed<br />

hairs; stipules 3-8 mm long, ovate-, lance- or deltoid-acuminate,<br />

mostly recurved, partially or fully amplexicaul-decurrent,<br />

none connate; leaves 6-16 cm long;<br />

leaflets 15-23 (25), ovate, suborbicular, or obovate, the<br />

apex obtuse or emarginate, 5-22 mm long; peduncles<br />

erect, 5-14 cm long; racemes 13-30 (35) flowered, early<br />

elongating, flowers ascending to spreading, the axis becoming<br />

(3) 5-12 (20) cm long in fruit; calyx 7-9 mm<br />

long, white-, black-strigulose or mixed, the campanulate<br />

or cylindric tube 4.5-5.5 (6.5) mm long, the teeth, subulate<br />

to lance-acuminate, 1-2.5 mm long; petals light<br />

lemon yellow, drying ochroleucous to brownish; banner<br />

14-16.5 mm long; keel (10) 11-13 mm long, immaculate;<br />

wings 12-15 mm long, whitish with light lemon<br />

yellow tip; ovary glabrous; ovules 22-26; fruiting pedicels<br />

persistent, ascending or spreading, straight or<br />

curved; pod persistent, ascending or spreading, straight<br />

or incurved less than 90, in longitudinal section linear or<br />

narrowly oblong, in cross section cordate or terete, (17)<br />

19-23 x 3-4 mm, 5-6 (6.2)x longer than wide, sessile on<br />

a minute boss on the receptacle or contracted at the base<br />

into an incipient stipe to 0.5 mm long, the valves thinly<br />

fleshy, becoming coriaceous, stramineous, semibiloculate<br />

to nearly biloculate (but not fused to the funicular<br />

flange), the septum 1.5-2 mm wide, the beak<br />

unilocular; dehiscence apical, through the beak while<br />

still attached to the raceme.<br />

Habitat. In mixed shrub communities, in sandy,<br />

gravely washes (sometimes among boulders) derived<br />

from Precambrian granites and Tertiary volcanic rocks.<br />

Distribution. In northern Maricopa County, found in<br />

the foothills and alluvial fans in vicinity of Cave Creek,<br />

Fish Creek, Scottsdale, and Tempe; to be expected in<br />

the foothills and alluvial fans from Scottsdale and<br />

Tempe east to the mountains along both sides of the<br />

Verde River drainage, southward to its confluence with<br />

the Salt River, and west to the alluvial fans in the vicinity<br />

of Saguaro Lake (see Lehto 510 from 1962 below).<br />

Phenology. Flowering from February - April; fruiting<br />

from April - June.<br />

Astragalus lentiginosus var. maricopae has been<br />

largely overlooked by most botanists since it was first<br />

described in 1945. Based on similar floral morphology,<br />

is has been confused with A. lentiginosus var. yuccanus.<br />

On the valley floor and alluvial fans northeast of Phoenix,<br />

Scottsdale, and Tempe, this taxon has become very<br />

rare (and nearly extirpated throughout its historically<br />

known range) due to extensive suburban housing and<br />

golf course development. The population sampled for<br />

molecular analysis in the vicinity of Scottsdale has already<br />

been developed, since home construction was<br />

well underway when the samples were collected. This<br />

variety is the most unique morphologically, and the<br />

most endangered of all the Palantia.<br />

Voucher specimens examined for the morphologic<br />

analysis. USA. ARIZONA: MARICOPA CO.:<br />

west of intersection of Westland Drive and Pima Rd,<br />

Scottsdale, February 2005 (fl, fr), Alexander 1621<br />

[individuals A, C, D, E] (OSC, UNLV); along Horseshoe<br />

Dam Rd, 0.5 mi below dam, 02 March 1989 (fl),<br />

C.L. Jones 5, (GH, NY, RSA)<br />

Voucher specimens examined (to be included in<br />

future morphological analyses). U.S.A. ARIZONA:<br />

148


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

MARICOPA CO.: 10 mi E of Scottsdale rd, 27 mi NE<br />

of Scottsdale, 24 March 1960 (fl, imm fr), Crosswhite et<br />

al 496 (NY); Scottsdale rd, between Bell Rd & Carefree,<br />

27 April 1974 (fl, fr), Engard et al. 203 (NY); 26<br />

mi NE of Scottsdale along Hwy 87, 28 March 1973 (fl),<br />

Higgins 6445 (NY); Hwy 87, 2.7 mi SW of Saguaro<br />

Lake, 14 April 1962 (fl, fr), Lehto 510 (NY); Cave<br />

Creek, 23 April 1977 (fl, fr), Lehto 21306 (NY); Carefree,<br />

23 April 1977 (fl, fr), Lehto 21308 (NY).<br />

2. Astragalus lentiginosus var. ursinus<br />

Astragalus lentiginosus var. ursinus (A. Gray) Barneby,<br />

Leafl. West. Bot. 4: 133. 1945. Astragalus ursinus A.<br />

Gray, Proc. Am. Acad. Arts Sci. 13: 367. 1878. Tium<br />

ursinum (A. Gray) Rydb., N. Amer. Fl. 24: 398. 1929.<br />

TYPE: U.S.A. ARIZONA OR UTAH: MOHAVE CO.<br />

OR WASHINGTON CO.: Beaver Dams [west slope of<br />

Beaver Dam Mountains, the type locality was erroneously<br />

published by Gray as Bear Valley, Iron Co.,<br />

<strong>Utah</strong>], 20-28 Apr 1877, E. Palmer s.n. (LECTOTYPE:<br />

GH, designated by Alexander, in prep; ISOLECTO-<br />

TYPE: K). Map: Figure 2.<br />

Perennial herbs, 2-4 dm tall; stems erect and ascending<br />

in clumps from a superficial root crown; herbage<br />

glabrous to sparsely strigulose with basifixed hairs; stipules<br />

3-8 mm long, triangular- or deltate-acuminate,<br />

mostly reflexed, partially or fully amplexicauldecurrent,<br />

none connate; leaves 2-9 cm long; leaflets 11-<br />

17 (19), suborbicular, obovate, or oblong, the apex obtuse<br />

or emarginate, 5-11 mm long; peduncles erect, 4-10<br />

cm long; racemes 7-15 (20) flowered, early elongating,<br />

flowers ascending, the axis becoming 3-9 (-11) cm long<br />

in fruit; calyx 4-8 mm long, white-, black-strigulose or<br />

mixed, the campanulate or cylindric tube 3-6 mm long,<br />

the teeth deltate, subulate to lance-acuminate, 0.8-3 mm<br />

long; petals pink purple, drying violet; banner 12-16<br />

mm long, purple with a white, purple striate spot; keel<br />

8.5-13 mm long, light to dark purple maculate; wings<br />

10.5-16 mm long, purple with dark purple tip or purple<br />

with a white tip; ovary glabrous or sparsely strigulose;<br />

ovules 22-24; fruiting pedicels persistent, erect or ascending,<br />

straight or curved; pod long-persistent, erect or<br />

ascending, in longitudinal section oblong or narrowly<br />

elliptic, in cross section cordate or terete, straight or incurved<br />

less than 90°, 10-23 x 4-5 mm, 2.4-4.7x longer<br />

than wide, sessile on a minute boss on the receptacle or<br />

contracted at the base into an incipient stipe to 0.7 (1.0)<br />

mm long, the valves thinly fleshy, becoming coriaceous,<br />

stramineous to reddish, semi-biloculate to nearly biloculate<br />

(but not fused to the funicular flange), the septum 2-<br />

2.5 mm wide, not extended into the beak, the ventral<br />

suture sometimes prominent, the beak unilocular; dehiscence<br />

apical, through the beak while still attached to the<br />

raceme.<br />

Habitat. In mixed shrub communities with Larrea and<br />

Yucca brevifolia, in gravely washes and talus slopes<br />

derived from the Permian Kaibab Formation (limestone),<br />

Toroweap Formation (limestones and sandstones),<br />

Hermit Formation (sandstones and siltstones),<br />

Queantoweap Sandstone, Permian-Pennsylvanian Callville<br />

Limestone, and Mississippian Redwall Limestone;<br />

with Penstemon petiolatus and other limestone crevice<br />

species on limestone cliffs of various Paleozoic limestones,<br />

especially the Kaibab Formation and Callville<br />

Limestone.<br />

Distribution. Washington Co., <strong>Utah</strong>, in the southern<br />

end of the Beaver Dam Mountains in the vicinity of<br />

Bulldog Knolls and Bulldog Canyon, north to Cedar<br />

Pockets Wash on the slopes of the peak south of Jarvis<br />

Peak; in adjacent Mohave Co., Arizona, south to the<br />

mouth of the Virgin River Gorge, and Hedricks Canyon<br />

in the Virgin Mountains; to be looked for in the vicinity<br />

of Mokaac Mountain, Wolf Hole Mountain, Quail Canyon<br />

or Quail Hill on the northern edge of the Shivwits<br />

Plateau, Mohave Co., Arizona.<br />

Phenology. Flowering from March - April; fruiting<br />

from April - May.<br />

The type collections of Astragalus ursinus are a<br />

drought depauperate, limestone crevice form of A. lentiginosus<br />

var. ursinus. The depauperate morphology of<br />

the types, especially with respect to the small flower<br />

size, has contributed to a perennial fog of confusion surrounding<br />

this variety's taxonomic relationships. It is<br />

only slightly differentiated morphologically from A.<br />

lentiginosus var. mokiacensis and imperfectly distinguished<br />

from some A. lentiginosus var. mokiacensis<br />

populations in habitat preference (with respect to the<br />

small number of plants per population growing within<br />

limestone crevices only). Both varieties have populations<br />

that inhabit limestone talus slopes below cliff<br />

faces. The genetic analysis (Alexander 2008, Alexander<br />

& Liston, in prep) shows that A. lentiginosus var. ursinus<br />

is distinct from A. lentiginosus var. mokiacensis and<br />

more closely related to A. lentiginosus var. palans and<br />

A. lentiginosus var. wilsonii. Additionally, Astragalus<br />

lentiginosus var. ursinus and A. lentiginosus var. wilsonii<br />

are the only two members of the Palantia with<br />

erect to ascending fruiting pedicels and erect to ascending,<br />

incurved to nearly straight pods.<br />

Of further note, Alexander (2005) cites the lectotypifications<br />

for Astragalus mokiacensis and A. ursinus as<br />

being published in Taxon. However, these two lectotypifications<br />

have not yet been published due to technical<br />

circumstances beyond the author's control. The citation<br />

of the lectotype above should not be considered the<br />

formal lectotypification of Astragalus ursinus.<br />

Voucher specimens examined for the morphological<br />

analysis. U.S.A. ARIZONA OR UTAH: MOHAVE<br />

CO. OR WASHINGTON CO.: Beaver Dams [Beaver<br />

149


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Dam Mountains, mislabeled as Bear Valley and Beaver<br />

Valley by Gray], 20-28 Apr 1877 (fl, fr), Palmer s.n.<br />

(GH, K). ARIZONA: MOHAVE CO.: Virgin Narrows<br />

about 5 mi NE of Littlefield, 2 Jun 1977 (fl, fr),<br />

Gierisch 3954 (BRY, NY); Hedricks Canyon [Virgin<br />

Mountains], T40N R15W S23, 2 Apr 1981 (fl, imm fr),<br />

Gierisch & Morgart 4824 (BRY, NY). UTAH: WASH-<br />

INGTON CO.: Beaver Dam Mountains, Bull Dog Canyon,<br />

25 Apr 2003 (fr), Alexander 1388 (OSC, UNLV), 6<br />

May 2005 (fl,fr), Alexander 2120 (OSC,UNLV), 6 May<br />

2005 (fl, fr), Alexander 2121 (OSC,UNLV), 6 May<br />

2005 (fl, fr), Alexander 2127 (OSC,UNLV); Bulldog<br />

Knolls, 6 May 2006 (fl, fr), Alexander 2132<br />

(OSC,UNLV), 6 May 2006, Alexander 2134<br />

[individuals A,B,C] (OSC, UNLV), 6 May 2006 (fl, fr),<br />

Alexander 2135 (OSC, UNLV); Bull Dog Knolls, S<br />

slope of S knoll, T43S R18W S28, 1097 m, 21 Apr<br />

1986 (fl, fr), Baird 2324 (BRY); Bulldog Knolls, north<br />

knoll, T43S R18W S21, 341 m, 28 Mar 1986 (fl, imm<br />

fr), Higgins & Barnum 16267 (BRY, NY); Bulldog<br />

Canyon, Beaver Dam Mountains, W slope, T43S R18W<br />

S26, 381 m, 15 Apr 1983 (fl, fr), Neese & Welsh 13062<br />

(BRY, NY); Cedar Pockets wash, along a sequential<br />

drainage, T43S R17W S20, 1219 m, 5 May 1986 (fr),<br />

Welsh & Atwood 23742 (BRY).<br />

ACKNOWLEDGMENTS<br />

Numerous individuals provided advice, funding and<br />

encouragement during various stages of this project including:<br />

Wesley E. Niles; Aaron Liston, my thesis advisor;<br />

Richard Halse of the OSC Herbarium; Kathryn<br />

Birgy of the UNLV Herbarium; Arnold Tiehm and the<br />

Nevada <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> (and the NNPS Small<br />

Grants Program); Lisa Decesare of the Library of the<br />

Gray Herbarium, Harvard University; the Moldenke<br />

Fund for <strong>Plant</strong> Systematics and the Hardman Fund at<br />

Oregon State University; and my committee members,<br />

Mary Santelmann, David Lytle, Rich Cronn, and Paul<br />

Farber. Additional thanks go to the late R.C. Barneby<br />

and the Barneby Fund for Research in Legume Systematics,<br />

which funded a month of research at NY. The<br />

following herbaria's online databases were consulted for<br />

this project: CAS, DS, GH, JEPS, K, MO, NY, OSC,<br />

ORE, RSA, UC, UNLV, US, WILLU. I thank the curators<br />

and staff of the following herbaria for access to<br />

their collections or loans of their specimens: BRY,<br />

CAS, DS, GH, JEPS, K, NY, OSC, ORE, RM, RSA,<br />

UC, UNLV, US, WILLU.<br />

LITERATURE CITED<br />

Alexander, J.A. 2005. The taxonomic status of Astragalus<br />

mokiacensis (Fabaceae), a species endemic to the<br />

Southwestern United States. Brittonia 57:320-333.<br />

Alexander, J.A. 2008. A taxonomic revision of Astragalus<br />

mokiacensis and allied taxa within the Astragalus<br />

lentiginosus complex of Section Diphysi. Ph.D.<br />

dissertation. Oregon State University, Corvallis, OR.<br />

Alexander, J.A. 2009. The types of Astragalus Section<br />

Diphysi (Fabaceae), a complex endemic to Western<br />

North America, Part I: lectotypifications, epitypifications,<br />

and new combinations of several taxa. Journal of<br />

the Botanical Research Institute of Texas 3:211-218.<br />

Barneby, R.C. 1945. Pugillus Astragalorum - IV: the<br />

section Diplocystium. Leaflets of Western Botany 4: 65-<br />

147.<br />

Barneby, R.C. 1956. Pugillus Astragalorum - XVII:<br />

four new species and one variety. Leaflets of Western<br />

Botany 8: 14-23.<br />

Barneby, R.C. 1964. Atlas of North American Astragalus.<br />

(2 volumes). Memoirs of the New York Botanical<br />

Garden. 13:1-1188.<br />

Barneby, R.C. 1989. Intermountain Flora. Volume<br />

III Part B. Eds. A. Cronquist, A.H. Holmgren, N.H.<br />

Holmgren, J.L. Reveal, P.K. Holmgren. The New York<br />

Botanical Garden. Bronx, New York. 279 pp.<br />

Cracraft, J. 1983. Species concepts and speciation<br />

analysis. Current Ornithology 1: 159-187.<br />

Easdale, T.A., Gurvich, D.E., Sersic, A.N. & Healey,<br />

J.R. 2007. Tree morphology in seasonally dry montane<br />

forest in Argentina: Relationships with shade tolerance<br />

and nutrient shortage. Journal of Vegetation Science 18:<br />

313-326.<br />

Gower, J.C. 1971. A General Coefficient of Similarity<br />

and Some of Its Properties. Biometrics 27:857-871.<br />

Gray, A. 1849. <strong>Plant</strong>ae Fendlerianae Novi-Mexicanae.<br />

Memoirs of the American Academy of Arts and<br />

Sciences. Series 2. 1:1-116.<br />

Gray, A. 1863. A Revision and Arrangement (mainly<br />

by the fruit) of the North American species of Astragalus<br />

and Oxytropis. Proceedings of the American Academy<br />

of Arts and Sciences. 6:188-237.<br />

Gray, A. 1865. Characters of some new plants of<br />

California and Nevada, chiefly from the collections of<br />

Professor William H. Brewer, botanist of the State Geological<br />

Survey of California, and of Dr. Charles L.<br />

Anderson, with revisions of certain genera or groups.<br />

Proceedings of the American Academy of Arts and Sciences.<br />

6:519-556.<br />

Gray, A. 1878. Contributions to the botany of North<br />

America. Proceedings of the American Academy of<br />

Arts and Sciences. 13:361-374.<br />

Hammer, O., Harper, D.A.T., and P.D. Ryan. 2001.<br />

PAST: Palaeontological Statistics software package for<br />

education and data analysis. Palaeontological Electronica<br />

4:9 pp.<br />

Hooker, W.J. 1831. Flora Borelli-American or the<br />

botany of the northern parts of British America. 2<br />

(12):97-160.<br />

150


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Isely, D. 1998. <strong>Native</strong> and naturalized Leguminosae<br />

(Fabaceae) of the United States. Brigham Young University.<br />

Provo, <strong>Utah</strong>. 1007 pp.<br />

Jones, M.E. 1898. Contributions to Western Botany<br />

8. Contributions to Western Botany. 8:1-43.<br />

Jones, M.E. 1923. Revision of North American species<br />

of Astragalus. Text distributed Feb. 15, 1923;<br />

Plates, June 20, 1923. Salt Lake City, <strong>Utah</strong>. 288 pp.<br />

Kearney, T.H. and R.H. Peebles. 1960. Arizona<br />

Flora. University of California Press. Berkeley, California.<br />

1079 pp.<br />

Knaus, B.J. 2008. A Fistful of Astragalus: phenotypic<br />

and genotypic basis of the most taxon rich species<br />

in the North American Flora. Ph.D. dissertation. Oregon<br />

State University, Corvallis, OR.<br />

Luckow, M. 1995. Species Concepts: assumptions,<br />

methods, and applications. Systematic Botany. 20:589-<br />

605.<br />

Munz, P.A. and D.D. Keck. 1959. A California flora.<br />

University of California Press. Berkeley, California.<br />

Nixon, K.C., and Q.D. Wheeler. 1990. An amplification<br />

of the phylogenetic species concept. Cladistics 6:<br />

211-223.<br />

Rydberg, P.A. 1929. Astragalinae. North American<br />

Flora 24(5): 251-462.<br />

Sokal, R. R. 1973. The species problem reconsidered.<br />

Systematic Zoology. 22: 360-374.<br />

Swofford, D. L. 2002. PAUP*: phylogenetic analysis<br />

using parsimony. ver. 4 beta 10. Sinauer Associates,<br />

Inc., Sunderland, Massachusetts.<br />

Watson, S. 1871. Botany. In C. King, Report of the<br />

geological exploration of the fortieth parallel, Vol. 5.<br />

Washington: Government Printing Office. 525 pp.<br />

Welsh, S.L. 1978. <strong>Utah</strong> flora: Fabaceae (Leguminosae).<br />

Great Basin Naturalist. 38:225-367<br />

Welsh, S.L. 1993. Leguminosae (Fabaceae). Pp 379-<br />

463 in S. L . Welsh, N. D. Atwood, S. Goodrich, and L.<br />

C. Higgins, eds., A <strong>Utah</strong> Flora. 2nd ed. Monte L. Bean<br />

Life Science Museum, Provo, <strong>Utah</strong>.<br />

Welsh, S.L. 2003. Leguminosae (Fabaceae). Pp 350-<br />

427 in S. L . Welsh, N. D. Atwood, S. Goodrich, and L.<br />

C. Higgins, eds., A <strong>Utah</strong> Flora. 3rd ed. Monte L. Bean<br />

Life Science Museum, Provo, <strong>Utah</strong>.<br />

Welsh, S.L. 2007. North American species of Astragalus<br />

(Leguminosae): A taxonomic review. Monte L.<br />

Bean Life Science Museum. Provo, <strong>Utah</strong>. 932 pp.<br />

Welsh, S.L., and N.D. Atwood. 2001. New taxa and<br />

nomenclatural proposals in miscellaneous families-<br />

<strong>Utah</strong> and Arizona. Rhodora. 103: 71-95.<br />

Wojciechowski, M. F., M. J. Sanderson, B. G. Baldwin,<br />

& M. J. Donoghue. 1993. Monophyly of aneuploid<br />

Astragalus (Fabaceae): evidence from nuclear ribosomal<br />

DNA internal transcribed spacer sequences. American<br />

Journal of Botany. 80: 711-722.<br />

Wojciechowski, M.F., M.J. Sanderson, J.-M. Hu.<br />

1999. Evidence on the monophyly of Astragalus<br />

(Fabaceae), and its major subgroups based on nuclear<br />

ribosomal DNA ITS and chloroplast DNA trnL intron<br />

data. Systematic Botany. 24: 409-437.<br />

APPENDIX 1<br />

List of vouchers used in morphometric analyses not<br />

cited in the Taxonomic Revision<br />

Note: All vouchers listed herein have been seen by<br />

the author. The "!" designation is not used herein as a<br />

result.<br />

Astragalus lentiginosus var. ambiguus Barneby<br />

U.S.A. ARIZONA: MOHAVE CO.: Peach Springs, 11<br />

May 1941, Ripley & Barneby 3403 (RSA).<br />

Astragalus lentiginosus var. araneosus (Sheld.)<br />

Barneby U.S.A. UTAH: MILLARD CO.: Frisco, June<br />

1880, Jones 1807 [June 1880, Jones s.n., interpreted<br />

herein as an isotype] (NY, ORE, GH).<br />

Astragalus lentiginosus var. bryantii (Barneby)<br />

J.A. Alexander U.S.A. ARIZONA: COCONINO CO.:<br />

head of Phantom Canyon in Grand Canyon, 15 <strong>December</strong><br />

1939, Bryant s.n. (CAS); 10 yds N of Colorado<br />

River, 11 mi S of Phantom Ranch, directly N of Grand<br />

Canyon Village, 11 April 1960, Crosswhite 642 (NY);<br />

at mouth of Hermit Creek, in sand, Grand Canyon of the<br />

Colorado River, 10 April 1917, Eastwood 5991 (GH);<br />

"<strong>Utah</strong> Flat", Grand Canyon N.P., ca. 0.83 mi NW of<br />

Phantom Ranch and Bright Angel Creek, 09 April 1993,<br />

Hodgson & Anderson 2085 (NY); Colorado River, Bass<br />

Rapids, 108 miles below Lees Ferry, 01 May 1971,<br />

Holmgren, et al. 15502 (NY); Colorado River, Grand<br />

Canyon near confluence of Clear Creek, 3.5 miles upriver<br />

from Kaibab Suspension Bridge (near Phantom<br />

Ranch), 1 mile up Clear Creek Canyon, 08 May 1971,<br />

Holmgren, et al. 15609 (GH, NY); Colorado River,<br />

Grand Canyon, Shinumo Creek, 108.5 river mi below<br />

Lees Ferry, 17 air mi NW of Grand Canyon Village, 10<br />

May 1971, Holmgren, et al. 15615 (NY); Grand Canyon<br />

N.P., at confluence of Bright Angel Creek and Colorado<br />

River, 20 March 1968, Spellenberg 1826.<br />

Astragalus lentiginosus var. iodanthus (S. Watson)<br />

J.A. Alexander U.S.A. NEVADA: CARSON CITY<br />

(ORMSBY) CO.: Empire City [Carson City vicinity],<br />

19 May 1882, Jones 3837 [May 1882, Jones s.n., interpreted<br />

herein as a duplicate] (NY, ORE).<br />

151


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Astragalus lentiginosus var. mokiacensis (A. Gray)<br />

M. E. Jones U.S.A. ARIZONA: MOHAVE CO.: Hidden<br />

Canyon, ca. 0.5 mi W of corral, 36.5366°N<br />

113.7388°W, 1159 m, 25 May 2002 (fr), Alexander<br />

1304 (OSC, UVSC), 02 May 2003 (fl, fr), Alexander<br />

1398 (NY, OSC, UNLV, UVSC); Hidden Canyon wash,<br />

36.5291°N 113.7283°W, 1189 m, 02 May 2003 (fl, fr),<br />

Alexander 1500 (OSC, UNLV); Whitmore Canyon, below<br />

Kinney Point, 02 May 2003 (fl, fr), Alexander 1502<br />

(OSC, UNLV, UVSC); on the N slope of Garnet Mountain<br />

[Iron Mountain], 20 Apr 1997 (fl, fr), Alexander et<br />

al. 846 (UNLV); 11 mi S of Mt. Trumbull village,<br />

Parashant (Trail) Canyon, 26 Apr 1974 (fl, fr), Atwood<br />

6029 (BRY, NY); 30 mi S of Mt. Trumbull village,<br />

Andrus Canyon, 28 Apr 1974 (fl, imm fr), Atwood 6056<br />

(NY); 9 mi S of Mt. Trumbull village, head of Parashant<br />

(Trail) Canyon, 28 Apr 1974 (fl, imm fr), Atwood 6087<br />

(BRY, NY); Andrus Canyon, 3 mi W of Andrus Point,<br />

T32N R10W S6, 26 Apr 1999 (fr), Atwood & Furniss<br />

24293 (BRY, NY, RM, UNLV, US); Andrus Canyon, 3<br />

mi W of Andrus Point, T32N R10W S6, 26 Apr 1999<br />

(fl, imm fr), Atwood & Furniss 24300 (BRY, NY);<br />

Andrus Canyon, 1 mi W of Andrus Point, T32N R10W<br />

S10, 26 Apr 1999 (fl, fr), Atwood & Furniss 24302<br />

(BRY, NY); drainage below Andrus Spring, T33N<br />

R12W S20, 19 Apr 2000 (fl, imm fr), Atwood et al.<br />

25058 (BRY, NY); 1 mi S of Trail Canyon summit,<br />

Parashant-Andrus rd, T33N R10W S2, 19 Apr 2000 (fl,<br />

fr), Atwood et al. 25095 (BRY, NY); 2 mi S of Mt.<br />

Trumbull school house, near Griffiths Knoll, T34N<br />

R10W S1-S2, 1600 m, 21 Apr 2000 (fl, fr), Higgins et<br />

al. 20277 (BRY, NY); Bar Ten Ranch [Hells Hollow],<br />

T33N R9W S14, 21 Apr 2000 (fl, fr), Higgins et al.<br />

21171 (BRY, NY, OSC); 12.5 km (7.8 miles) south of<br />

Mt. Trumbull, Whitmore Canyon, T34N R9W S29,<br />

1600 m, 25 May 1979 (fl, fr), Holmgren et al. 9172<br />

(BRY, NY); Shivwits Plateau, wash 0.5 mi W of Cupe<br />

Spring [Cupe Seep], Grassy Point, Lake Mead National<br />

Recreation Area, 21 May 1977 (fl, fr), Leary 1646<br />

(UNLV); Grand Canyon of the Colorado River near<br />

Peach Springs [erroneously labeled as “Fort Mohave”],<br />

May 1884 [Apr 27] (fr), Lemmon 3116 (GH); Peach<br />

Spring, on hills, Grand Canyon of the Colorado River,<br />

Jun 1884 (fr), Lemmon 3326 (GH, UC); Mokiak Pass,<br />

[Palmer's “Juniper Mountains” in the vicinity of Grand<br />

Wash], 28 Apr-2 May or 2-4 Jun 1877 (fl, fr), Palmer<br />

105 (GH, K, NY, MO, POM, US); top of Grand Wash<br />

Cliffs above Vulture Canyon [Andrus Canyon vicinity],<br />

at lowermost end of Grand Canyon, T32N R10W S6,<br />

1280 m, 19 Mar 1977 (fl, imm fr), Phillips III 77-1<br />

(NY); [Whitmore Canyon vicinity], T34N R9W S18,<br />

1585 m, 2 Jun 1978 (fl, fr), Smith & Gierisch 1091<br />

(BRY, NY, 2 sheets); Cottonwood Wash, sandy wash<br />

bottom, T37N R15W S28, 1539 m, 20 May 1987 (fl, fr),<br />

Thorne & Atwood 5256 BRY); NEVADA: CLARK<br />

CO.: Gold Butte, NW edge, Granite Spring vicinity,<br />

36.2847°N 114.1920°W, 10 Apr 2001 (fl, fr), Alexander<br />

1147 (OSC, UNLV); Gold Butte, southwest foothills,<br />

36.2712°N 114.2137° (W, 10 Apr 2001 (fl, fr), Alexander<br />

1148 (OSC, UNLV); Quail Springs Wash, Gold<br />

Butte area, 36.2585°N 114.2028°W, 1220 m, 3 May<br />

2003 (fl, fr), Alexander 1503 (OSC); Grapevine Spring,<br />

Gold Butte area, 36.23990°N 114.1742°W, 4200 ft<br />

(1280 m), 3 May 2003 (fl, fr), Alexander 1505 (OSC,<br />

UNLV); Twin Springs Wash, 36.1819°N 114.2222°W,<br />

976 m, 17 May 2003 (fl, fr), Alexander 1510 (OSC,<br />

UNLV); Black Mountains, northwest slope, on a ridge E<br />

of Pinto Valley, Lake Mead National Recreation Area,<br />

Z11 723894 m E 4012957 m N, 900 m, 13 Apr 1997 (fl,<br />

fr), Alexander & Birgy 795 (UNLV); E of Gold Butte,<br />

ca. 1.6 mi S of Summit Pass, in gravely wash, 31 May<br />

2001 (fr), Alexander & Carlson 1160 (OSC); Mica<br />

Spring [Gold Butte area], 1219 m, 13 Apr 1894 (fl, fr),<br />

Jones 5058 (BRY, NY, UC, US); 3.1 mi S of Gold<br />

Butte in Cataract Wash, 20 May 1977 (fl, imm fr),<br />

Leary & Niles 1910 (NY, UNLV); Grapevine Spring<br />

Creek, near Jumbo Peak, near reservoir, Gold Butte<br />

area, T19S R70E S3, 1280 m, 26 May 1977 (fl, fr),<br />

Leary & Niles 1947 (UNLV); Quail Spring Wash, about<br />

1.3 mi SW of Gold Butte, 25 Apr 1997 (fl, fr), Niles et<br />

al. 4806 (UNLV); Garden Spring area, 3 mi NE of Gold<br />

Butte, T19S R70E S11, 1150 m, 19 Apr 1997 (fl, fr),<br />

Niles et al. 4932 (NY, UNLV); Garnet Valley, north<br />

base of Bonelli Peak, T20S R69E S25, 9 May 1997 (fl,<br />

fr), Niles et al. 4988 (UNLV); New Spring Wash, ca 1<br />

rd mi S of Summit Pass, 1128 m, 15 Apr 1986 (fr), Pinzl<br />

7032 (NY, UNLV).<br />

Astragalus lentiginosus var. palans (M.E. Jones)<br />

M.E. Jones U.S.A. ARIZONA: COCONINO CO.: ca<br />

0.5 mi S of Page, 01 May 2000, Atwood & Welsh<br />

25360 (NY, RM); E abutment of Glen Canyon Dam, S<br />

of hwy, 5 May 1998, Atwood & Welsh 26924 (NY);<br />

Glen Canyon damsite, 7 June 1961, Barneby 13114<br />

(CAS, NY, RSA); ca 2 rd mi S of Page, along AZ 89, 11<br />

May 1991, Christy 493 (NY); Navajo Power <strong>Plant</strong> pipeline,<br />

4 mi SE of Page, 15 April 1972, Davey s.n. (NY);<br />

base of Leche-E Rock, 15 April 1973, Hevly & Jenness<br />

s.n. (NY); Lake Powell, NW of Page, ca 0.5 mi NE of<br />

Glen Canyon Dam, 3 May 1996, Hufford 1130 (NY);<br />

Bitter Springs on rd to Page, 19 May 1973, LeDoux et<br />

al. 753 (NY), LeDoux et al. 781 (NY); US Hwy 89, ca 8<br />

mi SW of Page, 22 May 1973, Spellenberg et al. 3228<br />

(NY); NE part of Antelope Island, ca 5 mi NNE of<br />

Page, 14 April 1987, Tuhy & Holland 2927 (NY); NA-<br />

VAJO CO.: 12 mi N of Kayenta, 6 June 1961, Barneby<br />

13104 (CAS, NY, RSA); Mystery Valley, in region of<br />

Monument Valley, 16 April 1963, McClintock s.n.<br />

(CAS), 11 April 1963, McClintock s.n. (CAS); 4 mi W<br />

of Kayenta, 30 April 1981, Welsh 20381 (NY, RSA);<br />

152


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

COLORADO: DELTA CO.: Dominguez Creek, W of<br />

the Gunnison River below Bridgeport, 0-1.5 mi up<br />

Dominguez Canyon, 19 May 1982, Atwood & Thompson<br />

8773 (NY); MESA CO.: NW end of Sinbad Valley<br />

at head of Salt Creek Canyon, 26 May 1983, Atwood<br />

9262 (CAS, NY); 2 mi E of Bedrock, Paradox Valley,<br />

23 May 1984, Atwood et al. 9728 (NY); 6 mi SSW of<br />

Grand Junction, Rough Canyon, base of sandstone wall,<br />

18 May 1988, Dorn 4886 (NY, RM); open slope 10 mi<br />

S of Gateway, 11 June 1949, Harrington 4423 (RM);<br />

Grand Junction, Colorado Monument Park, 03 June<br />

1921, Osterhout 6142 (RM, RSA); Colorado N.M., west<br />

entrance, white hills, 17 May 1982, Siplivinsky 3273<br />

(RM); Colorado N.M., head of Ute Canyon, within park<br />

boundary, 25 May 1982, Siplivinsky 3379 (RM); near<br />

hq. Colorado N.M., 6 mi S of Fruita, mesa summit, 21<br />

May 1948, Weber 3840 (CAS, DUD, RM, US); MON-<br />

TROSE CO.: La Sal Creek, 4 mi S of Paradox, near cliff<br />

dwellers mine, 20 May 1982, Atwood & Thompson<br />

8798 (NY, RM); Monogram Mesa, 6 mi W of Vancorum,<br />

3 June 1961, Barneby 13046 (CAS, NY, RSA);<br />

Paradox Valley, N slope, 25 April 1986, Franklin 2838<br />

(NY, RM); ca 18 air mi NW of Naturita, ca 2.5 air mi S<br />

of Bedrock, E side Dolores River Canyon, E facing<br />

slope on W side canyon, 17 June 1995, Moore 5502<br />

(RM); Slick Rock Canyon, Dolores River, ca 18 air mi<br />

W of Naturita, at mouth and surrounding area of Leach<br />

Creek, SW facing wash, 05 July 1995, Moore 6334<br />

(RM); Slick Rock Canyon, Dolores River, ca 18 air mi<br />

W of Naturita, at mouth and surrounding area of Leach<br />

Creek, SW facing wash, 05 July 1995, Moore 6335<br />

(RM); near Naturita, SW Colorado, dry hills, 22 May<br />

1914, Payson 335 (RM); SAN MIGUEL CO.: Colorado<br />

side of Island Mesa, 27 May 1998, Atwood & Trotter<br />

23617 (NY); roadside, rocky cedar breaks,1 mi S of<br />

Gladel, 09 June 1951, Penland & Hartwell 4178 (RM);<br />

W end Gypsum Valley, rocky mouth of Hamm Canyon,<br />

09 June 1949, Weber 4735 (CAS, DUD, RM, RSA, UC,<br />

US); UTAH: CARBON CO.: ca 1.5 mi N of Emery Co.<br />

line along US Hwy 50-6, 29 April 1965, Welsh 3875<br />

(NY); EMERY CO.: San Rafael Swell, Chimney Rock<br />

flats, 17 May 1979, Harris 116 (RM); near junction of<br />

Lower Black Box Rd to Swazys Leap and Sulphur<br />

Spring, 19 May 1992, Heil & Hyder 7132 (RSA); San<br />

Rafael Swell, near Temple Mtn, 30 April 1968, Higgins<br />

& Reveal 1277 (NY, RM); San Rafael Swell, 18 May<br />

1915, Jones s.n. (UC), 12 May 1914, Jones s.n. (POM);<br />

Red Plateau, SW of Woodside, along rd to Castledale,<br />

6.9 mi W of US 50, 5 June 1958, Raven 13079 (CAS,<br />

GH, NY); Red Plateau, SE of Woodside, 13 June 1947,<br />

Ripley & Barneby 8675 (RSA); rocky draw, San Rafael<br />

Swell, 22 mi W of Green River on I-70, 1mi E of rest<br />

area just below Rattlesnake Bench, 28 April 1979,<br />

Shultz et al. 3113 (NY, RM, RSA); Summerville<br />

[Wash], at Woodside, 27 April 1977, Welsh & Taylor<br />

14624 (NY); GARFIELD CO.: upper E end of Choprock<br />

Bench, ca 30 mi ESE of Escalante, 6 May 1987,<br />

Tuhy & Holland 3127 (NY, RSA); head of North Wash,<br />

28 April 1981, Welsh 20368 (NY, RSA); GRAND CO.:<br />

Kane Springs Rd, 13 mi S of Moab, [no date], Atwood<br />

7478 (NY); Sandflat Rd, E of Moab 12 mi., ca 0.25 mi<br />

W of Forest Service boundary, 20 May 1982, Atwood &<br />

Thompson 8787 (NY,RM); Arches N.M., 5 mi NW of<br />

Moab, 21 May 1984, Atwood et al. 9695 (NY); 11 mi N<br />

of Moab, 18 May 1955, Barneby 12753 (CAS, NY,<br />

RSA); 1 mi N of Moab, 14 April 1940, Beath 48 (RM);<br />

Arches road, 04 May 1947, Beath s.n. (RM); near<br />

Moab, 17 May 1940, Beath & Goodding 6-373 (RM); 8<br />

mi NW of Moab, along UT Hwy 160, 28 April 1961,<br />

Bright 135 (NY); between Moab and bridge, 13 May<br />

1933, Cottam 5623 (RM); hillside in the Colorado River<br />

Canyon, a little below Salt Wash, NE of Moab, 9 May<br />

1961, Cronquist 8974 (GH, NY, RSA); slope E above<br />

Hwy 128, 0.15 mi N of Pole Canyon, 05 May 1985,<br />

Franklin 1391 (NY, RM); Red Hills SE above White<br />

Ranch along Colorado River, 07 May 1985, Franklin<br />

1421 (NY, RM); 3.7 mi due E of Moab, sand flats<br />

among sandstone fins, 17 April 1986, Franklin 2679<br />

(RM); Mat Martin Point, sand pockets on slickrock, 12<br />

May 1986, Franklin 2933 (RM); Sevenmile Mesa, point<br />

W of confluence of Dolores River and Fisher Cr., ca 25<br />

mi due NE of Moab, 18 May 1986, Franklin 3104<br />

(RM); Onion Creek, 22 May 1984, Goodrich et al.<br />

20398 (NY); N end of La Sal Mountains, Fisher Valley,<br />

22 May 1984, Goodrich et al. 20414 (NY); 1 mi NE of<br />

Moab, 0.5 mi along UT Hwy 128, along Colorado<br />

River, 30 April 1961, Hanson 148 (NY); on edge of<br />

Colorado River, Moab, clay buttes, 09 May 1933, Harrison<br />

5945 (RM); cliffs above headquarters, S edge of<br />

Arches N.M., sandy terraces among rocks, 25 April<br />

1947, Harrison 11123 (RSA, US), Harrison s.n. (UC);<br />

south of Courthouse Towers, Arches N.M., sandy flat,<br />

17 May 1950, Harrison 11383 (US, UC); The 'Neck" 13<br />

miles due WSW of Moab, sand rock ridge, 11 June<br />

1941, Harrison et al. 10286 (US); Arches N.M., 29 June<br />

1948, Howell 24752 (CAS,RSA); near Skyline Arch,<br />

Arches N.M., 7 September 1968, Howell & True 44906<br />

(CAS,NY); 10 mi N of Moab, sandy plain, 17 June<br />

1955, Isely 6460 (RSA, US); Castle Valley, May 1931,<br />

Jones s.n. (POM); red sandstone banks just north of<br />

Kane Springs, ca. 15 miles south of Moab, 21 April<br />

1966, Ledingham 4694 (UC); 5 mi N of Moab, 08 June<br />

1939, Porter 1796 (RM), Porter 1797 (RM); San Sige<br />

[illegible] Hollow, Grand River Canon [Grand River<br />

Canyon near Moab, June 1899], 1899, Purpus s.n. (UC);<br />

sandy slopes near headquarters, Arches N.M., near<br />

Moab, 6 June 1962, Rever & Belcher 73 (NY); W of<br />

Thompson, 13 June 1947, Ripley & Barneby 8662<br />

(RSA); Spanish Valley, 4.5 mi E of Moab, along roadside<br />

in town, 03 May 2000, Spencer 1487 (NY, RM);<br />

153


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Ida Gulch, 2.5 mi SE of Hwy 128 along Colorado River,<br />

E of Castle Valley Rd, slopes below Priests and Nuns,<br />

11 May 1986, Thorne et al. 4614 (RM); Sego, 2 mi N of<br />

Thompson, 30 April 1965, Welsh 3881 (NY); Castle<br />

Valley, ca 4.5 mi E of junction with UT Hwy 128, 3<br />

May 1968, Welsh 7018 (NY); Castle Valley, ca 4.5 mi E<br />

of junction with UT Hwy 128, 29 May 1968, Welsh &<br />

Moore 7170 (NY); KANE CO.: 5 mi S of Glen Canyon<br />

N.R.A. boundary, along Hole-in-the-Rock Rd., 23 April<br />

1996, Atwood & Furniss 20795 (NY); Willow Tank, 41<br />

mi SE of Escalante, 8 June 1961, Barneby 13127 (CAS,<br />

NY, RSA); abandoned "Pareah" townsite near the Paria<br />

River, 3 May 1986, Shultz & Shultz 9931 (NY); Willow<br />

Tank, ca 17 mi S of Garfield Co. line, along road to<br />

Hole-in-the-Rock, 04 May 1962, Welsh 1687 (NY, US);<br />

Lake Powell, Driftwood Canyon, Driftwood Garden, 27<br />

May 1986, Welsh 22102 (NY); Paria townsite, 5 May<br />

1970, Welsh & Atwood 9727 (NY); SAN JUAN CO.:<br />

south of Mexican Hat along Hwy 63, 14 May 1970, Atwood<br />

2486 (NY); Navajo Twins, N side of Bluff, 20<br />

May 1985, Atwood & Furniss 11009 (NY); E of Monticello,<br />

28 April 1998, Atwood & Trotter 23411 (NY); ca<br />

15 mi W of Mexican Hat, along bench above San Juan<br />

River, 28 April 1998, Atwood & Trotter 23423 (NY); E<br />

slope of Navajo Mountain, 23 June 1973, Atwood &<br />

Trotter 5334 (NY); Copper Canyon, S of Monitor Mesa,<br />

22 May 1985, Atwood et al. 11057 (NY); 14 mi W of<br />

Blanding, 19 May 1955, Barneby 12785 (CAS, NY,<br />

RSA); 10 mi W of Blanding on US Hwy 95, 30 April<br />

1961, Bright 83 (NY); Goose Necks, 01 May 1935, Cottam<br />

5837 (RM); between Blanding and Bluff, hillsides,<br />

17 April 1928, Cottam 6693 (RM, 2 sheets), Cottam<br />

6703 (RM); Goosenecks of San Juan River, 12 April<br />

1938, Cronquist 1104 (RM); Montezuma Canyon, 29<br />

May 1892, Eastwood s.n. (GH); Montezuma Canon, 01<br />

June 1892, Eastwood s.n. (RM, POM, UC); 25 mi W of<br />

Hanksville, 14 May 1940, Goodding & Beath 48 (RM);<br />

8 mi E of Halls Crossing, 23 May 1983, Higgins &<br />

Welsh 13216 (NY); ca 5 mi E of Mexican Hat, 24 May<br />

1983, Higgins & Welsh 13291 (NY); slopes above<br />

Gretchen Bar, 54 mi below Hite on the Colorado River,<br />

04 May 1954, Holmgren & Goddard 9980 (CAS); black<br />

streak cliffs, vicinity of Bluff, 24 June 1944, Holmgren<br />

& Hansen 3439 (NY); Mendenhall Loop, riverside below<br />

the Mendenhall Stone Cabin, ca mi 31, 2.5 km W of<br />

Mexican Hat, 11 June 1997, Holmgren & Holmgren<br />

12796 (NY); Bluff, 24 May 1919, Jones s.n. (POM); 14<br />

mi due SW of La Sal summit of Rone Bailey Mesa, 4<br />

June 1985, Neese & Welsh 16993 (NY); Bluff, south<br />

exposure, 03 April 1966, Pederson 12 (GH); W of<br />

Bluff, 22 May 1943, Ripley & Barneby 5603 (RSA);<br />

along San Juan River, near Bluffs, 25-29 August 1911,<br />

Rydberg & Garrett 9914 (NY, RM, US); 12.2 mi S of<br />

Moab, 22 May 1976, Shultz 2151 (RSA), Shultz &<br />

Bolander 2151 (NY); Bluff area, 1 mi E of Needles<br />

Overlook, along roadsides in sand, 22 May 1976, Shultz<br />

& Shultz 1854 (NY,US); Rone Bailey Mesa ca 22 mi<br />

NNW of Monticello, 01 July 1984, Tuhy 1581 (RM);<br />

head of Red Canyon, ca 12 mi SW of Natural Bridges<br />

N.M., 29 April 1981, Welsh 20374 (RSA); Bluff, 29<br />

April 1961, Welsh 1501 (NY); Comb Reef, 17 mi W of<br />

Blanding, 1 mi from Perkins Ranch, 30 April 1961,<br />

Welsh 1510 (NY); head of Red Canyon, ca 12 mi SW of<br />

Natural Bridges N.M., 29 April 1981, Welsh 20374<br />

(NY); Whirlwind Draw, Clay Hills Divide, 30 April<br />

1966, Welsh 5206 (NY); W of Grand View Point, top of<br />

Murphy's Ridge, 18 May 1968, Welsh 7076 (NY); S Six<br />

Shooter Peak platform, Davis Canyon, 11 May 1982,<br />

Welsh et al. 21100 (NY); between Natural Bridges N.M.<br />

and Blanding along UT Hwy 95, 22 April 1967, Wetherell<br />

& Finzel 633 (NY); WASHINGTON CO.: 2 mi W<br />

of Springdale, 15 April 1938, Cronquist & Gaupin 1106<br />

(BRY); Hurricane, 4 mi due SW of Rockville, 22 May<br />

1977, Foster 3766 (NY); Coalpits Wash, 5 May 1988,<br />

Franklin & Thorne 5930 (BRY); 9 mi E of Hurricane<br />

along Hwy 59, 20 May 1971, Higgins & Welsh 4227<br />

(NY); sandy desert, 5 miles east of Virgin, 11 May<br />

1937, Hitchcock 3025 (GH); Virgin, 14 May 1894,<br />

Jones 5215e (US); Rockville, 14 May 1894, Jones<br />

5215e (DUD, 3 sheets, NY, UC); Rockville, 15 May<br />

1894, Jones 5218 (DUD, GH, NY, RM), 14 May 1894,<br />

Jones 5218 (RM, US, UC); Virgen City, 14 May 1894,<br />

Jones 5218a (US); Rockville, 15 May 1894, Jones<br />

5218a (DUD, POM); Zion N.P, Petrified Forest, 13<br />

April 1972, R.A. Nelson 9930 (RM); Zion N.P, Petrified<br />

Forest, 13 April 1972, R.A. Nelson 9937 (RM); 2 mi N<br />

of Hurricane, 10 April 1966, Stevens 140 (BRY); Petrified<br />

Forest sector, 2 mi N of Rockville, 06 May 1988,<br />

Welsh & Clark 23986 (BRY, RM); WAYNE CO.: Capitol<br />

Reef entrance to Grand Wash, 24 May 1975, Harrison<br />

1688 (RM); Horseshoe Canyon, ca 1 mi up Horseshoe<br />

Canyon, Canyonlands N.P., 20 May 1990, Heil et<br />

al. 5895 (RSA); Capitol Reef entrance to Grand Wash,<br />

24 May 1975, K. Harrison 1688 (NY); canyon bed of<br />

Fremont Canyon, N of Fruita, 05 May 1940, Maguire<br />

18115 (RM); 23.8 mi W of Loa, 3 May 1977, Neese &<br />

White 2752 (NY); Barrier (Horseshoe) Canyon, 19 April<br />

1970, Welsh 9593 (NY); road to North Point ca 1 mi NE<br />

of French Spring above Orange Cliffs, 30 May 1970,<br />

Welsh & Atwood 9870 (NY).<br />

The following voucher specimens examined for the<br />

morphological analysis from the vicinity of Cameron<br />

are putative populations of A. lentiginosus var. palans<br />

intermediate to A. lentiginosus var. wilsonii.<br />

U.S.A. ARIZONA: COCONINO CO.: roadside, 6 mi W<br />

of Cameron, 08 April 1938, Cronquist & Gaupin 1105<br />

(RM); Painted Desert, Cameron, low areas and roadsides,<br />

22 April 1961, Demaree 43807 (DUD, NY, US);<br />

ca 5 mi W of Cameron, 8 April 1978, Gierisch 4183<br />

154


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

(NY); 4 mi W of Cameron on Hwy 64, 18 March 1968,<br />

Hitchcock 25614 (NY, OSC); Rainbow Lodge, south<br />

end of Navajo Mountain, 11 June 1938, Peebles &<br />

Smith 13939 (GH, NY, US); 10 mi SW of Cameron,<br />

Hwy 64, 7 April 1957, Strickland 21 (NY).<br />

Astragalus lentiginosus var. pseudiodanthus<br />

(Barneby) J.A. Alexander U.S.A. NEVADA: NYE<br />

CO.: sand dunes N of Tonopah, May 2003 (fl,fr), Alexander<br />

1631 (OSC, UNLV).<br />

Astragalus lentiginosus var. stramineus (Rydb.)<br />

Barneby U.S.A. ARIZONA: MOHAVE CO.:<br />

"southeastern <strong>Utah</strong>" [southwestern slope Beaver Dam<br />

Mountains, near Littlefield], June 1870, Palmer s.n.<br />

(NY, US).<br />

Astragalus lentiginosus var. vitreus Barneby<br />

U.S.A. UTAH: WASHINGTON CO.: 5 mi W of Leeds,<br />

19 May 1933, Maguire & Blood 4413. (POM, CAS).<br />

Astragalus lentiginosus var. wilsonii (Greene)<br />

Barneby U.S.A. ARIZONA: YAVAPAI CO.: sandy<br />

washes north of Cottonwood, 1 June 2005 (fl, fr), Alexander<br />

2334 [individuals B, E, J, K] (NY, OSC, UNLV,<br />

UVSC); south of the boundary of Montezuma Castle<br />

National Monument, 2 June 2005 (fr), Alexander 2339<br />

[individual F] (OSC, UNLV), 2 June 2005 (fl, fr), Alexander<br />

2340 [individuals A, D]; in scrubland 1.3 mi west<br />

of I-17, near junction of FR 9204F and AZ Hwy 179 2<br />

June 2005 (fr), Alexander 2367 [individuals A, B] (NY,<br />

OSC, UNLV, UVSC); on roadcut 0.1 mi west of I-17,<br />

along AZ Hwy 179, 02 June 2005 (fr), Alexander 2368<br />

(OSC, UNLV).<br />

The following vouchers specimens examined for the<br />

morphological analysis are putative populations of A.<br />

lentiginosus var. wilsonii along the South Rim of the<br />

Grand Canyon that are intermediate to A. lentiginosus<br />

var. bryantii, A. lentiginosus var. mokiacensis, or A. lentiginosus<br />

var. ursinus. U.S.A. COCONINO CO.: Kaibab<br />

Trail, 10 May 1940 (fl, fr), Collom KT24 (US);<br />

Grand View Trail, Grand Canyon of the Colorado<br />

River, 16 June 1916 (fl, fr), Eastwood 5748 (CAS, GH,<br />

UC); Grand Canyon [Bright Angel Trail vicinity], June<br />

1915 (fr), Macbride & Payson 945 (GH, RM); Kaibab<br />

Trail, Grand Canyon N.P., 23 May 1938 (fl, fr), Nelson<br />

& Nelson 2799 (RM); near El Tovar, South Rim Grand<br />

Canyon, 3 June 1947 (fr), Ripley & Barneby 8472<br />

(RSA).<br />

Astragalus lentiginosus var. yuccanus M.E. Jones<br />

U.S.A. ARIZONA: MOHAVE CO. : Yucca, 13 May<br />

1884, Jones 3886 (POM, GH, NY, US).<br />

Appendix II<br />

List of morphological characters and character<br />

states used in the morphometric analyses<br />

Adaxial leaflet pubescence (leafad): 1. densely<br />

pubescent (surface obscured, overlap); 2. moderately<br />

pubescent (some hair overlap, gap less than 0.2 mm); 3.<br />

sparsely pubescent to subglabrate (uneven to evenly<br />

across surface with no overlap, or just confined to midrib<br />

& base); 4. entirely glabrous (or just a few hairs at<br />

base)<br />

Abaxial (lower) leaflet pubescence (leafab): 1.<br />

densely pubescent (surface obscured, overlap); 2. moderately<br />

pubescent (some hair overlap, gap less than 0.2<br />

mm); 3. sparsely pubescent to subglabrate (uneven to<br />

evenly across surface with no overlap, or just confined<br />

to midrib & base); 4. entirely glabrous (or just a few<br />

hairs at base)<br />

Leaf and stem hair length (leafh): average of 3, in<br />

mm.<br />

Leaflet number (leafn): average of 3, in mm.<br />

Inflorescence length in flower (inflw): average of<br />

3, in mm.<br />

Calyx tube length (calyxl): average of 3, in mm.<br />

Calyx pubescence density (calyxd): 1. densely<br />

strigulose (surface obscured, overlap); 2. moderately<br />

strigulose (some hair overlap); 3. evenly & sparsely<br />

strigulose; 4. entirely glabrous (or just a few scattered<br />

hairs).<br />

Calyx teeth shape (calyxs): 1. deltoid; 2. lancesubulate,<br />

subulate; 3. subulate-setaceous; 4. lanceattentuate;<br />

5. lance-acuminate.<br />

Calyx teeth orientation (calyxo): 1. erect to spreading<br />

; 2. loosely long recurving.<br />

Keel length (keell): average of 3, in mm.<br />

Keel color (keelc): 1. light to dark purple maculate;<br />

2. tan to pink maculate tipped; 3. ochroleucous, not<br />

maculate; 4. yellow.<br />

Wing color (wingc): 1. purple, dark purple tipped; 2.<br />

purple, white tipped; 3. white, pink-purple tipped; 4.<br />

ochroleucous; 5. yellow.<br />

Banner color (bannc): 1. purple & white with purple<br />

striate central spot; 2. whitish to ochroleucous &<br />

pink tinged; 3. tan ochroleucous to whitish; 4. yellow.<br />

Inflorescence length in fruit (infr): average of 3, in<br />

mm.<br />

Pod pedicel orientation (podpo): 1. ascending to<br />

spreading, straight; 2. ascending to spreading, arched; 3.<br />

appressed to erect, straight; 4. recurved, arched<br />

Pod length & width ratio (podr): average of 3, in<br />

mm.<br />

Pod deciduous or persistent (poddp): 1. deciduous;<br />

2. persistent.<br />

155


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Pod shape, longitudinal section (podsl): 1. seculate<br />

& abruptly acute apically; 2. seculate & attenuate apically;<br />

3. elliptic, narrowly elliptic, lunate & acute apically;<br />

4. oblong, narrowly oblong & attenuate apically;<br />

5. linear; 6. lanceolate, widest at base & attenuate apically;<br />

7. oval, ovoid, broadly ovoid; 8. obovoid, broadly<br />

obovoid; 9. subglobose; 10. clavate-oblanceolate.<br />

Pod shape, cross section (podsc): 1. subterete, dorsiventrally<br />

compressed, when obcordate, the suture shallowly<br />

sulcate; 2. obcordate & laterally compressed,<br />

broadly to narrowly (obcompressed), suture deeply sulcate;<br />

3. didymous & dorsi-ventrally compressed,<br />

broadly to narrowly; 4. cordate & dorsi-ventrally compressed,<br />

broadly to narrowly; 5. cordate & laterally<br />

compressed, broadly to narrowly.<br />

Pod orientation on raceme (podro): 1. ascending,<br />

erect & beak & 1/2 pod curved inward; 2. ascending to<br />

spreading, straight; 3. spreading, declined & beak to 1/2<br />

pod recurved; 4. spreading & slightly curved inward<br />

from middle (lunate, falcate); 5. spreading & beak to 1/2<br />

pod incurved to 180° (hamate).<br />

Pod orientation, degree of pod incurve, coded<br />

range (podpi): 1.


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Ecology of Rusby’s Milkvetch (Astragalus rusbyi),<br />

a Rare Endemic of Northern Arizona Ponderosa Pine Forests<br />

Judith D. Springer, Michael T. Stoddard, and Daniel C. Laughlin,<br />

Ecological Restoration Institute, Northern Arizona University, Flagstaff, AZ<br />

Debra L. Crisp and Barbara G. Phillips,<br />

Coconino National Forest, Flagstaff, AZ<br />

Abstract. Rusby’s milkvetch (Astragalus rusbyi Greene) is endemic to basaltic soils northwest and west of Flagstaff,<br />

Arizona. Recent interest in this species is due in part to its addition to the U.S. Forest Service Region 3 sensitive species<br />

list in 1999 and its occurrence in ecological restoration projects and proposed fuels reduction projects that involve<br />

tree thinning and prescribed burning. Some of its habitat has been subjected to large wildfires over the last few<br />

decades, and other areas have undergone ecological restoration treatments, while much of its range in ponderosa pine<br />

forest is slated to undergo such treatments in the near future. In a ponderosa pine restoration study area northwest of<br />

Flagstaff, A. rusbyi was an indicator species of remnant grass patches and increased following tree thinning and prescribed<br />

burning. However, in an area less than 3 km away, there appeared to be no relationship to restoration treatments,<br />

trees per ha, pine basal area, or canopy cover, but A. rusbyi did appear to be sensitive to an extreme drought<br />

event in 2002 and may have remained dormant in that year, a pattern that has been observed in other rare Astragalus<br />

species. A. rusbyi has a foliar nitrogen content of 4.4% and a foliar C:N mass ratio of 11. It is classified as a competitive<br />

ruderal species, meaning it is able to compete well with other understory species, but is not very tolerant of<br />

stresses, such as deep shade. We currently do not have a thorough understanding of the ecology of this species, or the<br />

effects of ecological restoration or fuels reduction treatments. In this paper we will discuss ecology of other members<br />

of the genus Astragalus and explore the relationships of A. rusbyi to moisture, vegetation treatments and overstory<br />

mortality.<br />

Astragalus is believed to be the largest genus of<br />

flowering plants in the world, with over 2500 species<br />

worldwide and over 400 species in North America<br />

alone, primarily in arid regions of the western U.S. The<br />

highest diversity in North America is centered in the<br />

Great Basin and on the Colorado Plateau (Barneby<br />

1989, Sanderson 1991). Astragalus species are often<br />

found in marginal habitats or on specialized soil types,<br />

and their geographic ranges are strongly skewed towards<br />

narrow endemism (Barneby 1964, 1989; Sanderson<br />

1991). Along with a limited dispersal ability possessed<br />

by many members of the genus, there is widespread<br />

local differentiation and geographic speciation,<br />

particularly in the areas of the western U.S. where it<br />

achieves the highest levels of diversity (Sanderson<br />

1991, Lesica et al. 2006). However, due to restricted<br />

ranges and habitats, some Astragalus species may exhibit<br />

low genetic variability and reduced fitness from<br />

inbreeding depression (Karron et al. 1988, Allphin et al.<br />

2005, Breinholt et al. 2009). Neoendemism is common<br />

in the intermountain regions of North America where<br />

there are large numbers of both widespread, recently<br />

evolved species, as well as narrowly endemic species,<br />

which are often associated with extreme edaphic conditions<br />

and reduced competition from dominant species<br />

(Lesica et al. 2006). Lesica and his co-authors suggest<br />

that restricted ranges and high local abundances of<br />

neoendemic species may be due more to patterns and<br />

processes of speciation than to ecological tolerance. The<br />

small ranges exhibited by many Astragalus species in<br />

the western U.S. may be due to recent speciation and an<br />

insufficient amount of time for these species to have<br />

increased their ranges significantly (Lesica et al. 2006).<br />

Reticulate evolution may not be widespread in the Astragalus<br />

genus, for many members of the genus appear<br />

to exhibit allopatry (geographic isolation) along with<br />

high levels of local endemism and little hybridization<br />

(Sanderson 1991).<br />

The type specimen of A. rusbyi was collected by<br />

Henry Hurd Rusby on July 2, 1883 on Mt. Humphreys,<br />

near Flagstaff Arizona (Welsh 2007) and was first described<br />

by Edward Lee Greene in 1884 (Greene 1884).<br />

A. rusbyi is a slender perennial averaging 15-40 cm in<br />

height. It also has a fairly deep taproot (D.C. Laughlin,<br />

personal communication, 2008). It grows primarily in<br />

meadows in ponderosa pine forests and in aspen groves<br />

(Barneby 1964, Welsh 2007), but it also may be found<br />

in moderately dense ponderosa pine forests. Populations<br />

are mainly concentrated on basaltic soils in two areas in<br />

northern Arizona: around the San Francisco Peaks<br />

157


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

(primarily on the southern and western sides of the<br />

Peaks) and also the vicinity of Kendrick Mountain<br />

(Figure 1). Collections from areas to the south, including<br />

Oak Creek Canyon and Yavapai County are of uncertain<br />

validity. It is ranked G3 (vulnerable) by Nature-<br />

Serve (2009) and is on the U.S. Forest Service sensitive<br />

species list for Region 3 (Southwestern Region).<br />

A. rusbyi has been placed in the section Strigulosi,<br />

which contains approximately 35 species found mainly<br />

in the Mexican highlands north to Arizona and New<br />

Mexico and generally associated with oak and pine forests<br />

(Barneby 1964, Spellenberg 1974). Section Strigulosi<br />

is thought by some authors to be the most primitive<br />

group of Astragalus species in North America. The majority<br />

of the evidence, including research on chromosome<br />

numbers and more recent molecular phylogenetic<br />

data, currently points to an Old World origin for Astragalus,<br />

presumably in the steppes and mountains of<br />

southwestern and south-central Asia and the Himalayan<br />

Plateau (Spellenberg 1976, Wojciechowski 2005).<br />

Astragalus rusbyi has a chromosome number of 11<br />

(2n=22) (Spellenberg 1974). Morphologically, it appears<br />

to be most closely related to two other members of<br />

section Strigulosi, A. egglestonii and A. longissimus,<br />

with which it shares technical features (Barneby 1964).<br />

Like most members of Strigulosi, A. rusbyi flowers<br />

from mid-summer onward into the fall, varying in abundance<br />

in response to the amount and timing of summer<br />

rains. Astragalus rusbyi is differentiated from these<br />

species of Astragalus by minor differences in characters<br />

of its pendulous, stipitate, bilocular, trigonously compressed<br />

pods (Barneby 1964). However, the ranges of<br />

these three species do not overlap and they maintain<br />

geographic isolation from one another with no observed<br />

intermediate populations.<br />

ECOLOGY OF ENDEMIC WESTERN SPECIES<br />

Astragalus is a very large genus with little ecological<br />

information available for the vast majority of species.<br />

However, many uncommon and rare species share characteristics<br />

with each other that may directly affect monitoring<br />

and conservation planning for A. rusbyi. For example,<br />

some species exhibit prolonged vegetative dormancy,<br />

such as A. scaphoides (Bitterroot milkvetch) and<br />

A. sinuatus (Whited’s milkvetch) of the sagebrush<br />

steppe of the Pacific Northwest, or dormancy during dry<br />

years, as demonstrated by A. schmolliae (Schmoll’s<br />

milkvetch) of Mesa Verde National Park, in southwestern<br />

Colorado (Anderson 2004, Gamon 1995, Lesica and<br />

Steele 1994). A. scaphoides plants may utilize prolonged<br />

vegetative dormancy as a bet-hedging strategy in<br />

an effort to conserve resources while avoiding the risk<br />

inherent in funneling resources into aboveground<br />

growth (Gremer et al. <strong>2012</strong>). Lesica (1995) found that<br />

A. scaphoides plants may remain dormant belowground<br />

for as long as five years before reappearing. Dormancy<br />

may be inferred in other species, such as A. ripleyi, due<br />

to an increased number of visible plants in years of<br />

above average precipitation (Ladyman 2003). Barneby<br />

(1964) notes that A. rusbyi, like most members of section<br />

Strigulosi varies in “vigor and abundance in proportion<br />

to amount and timing of summer rains,” but prolonged<br />

vegetative dormancy has not been established.<br />

Many Astragalus species exhibit large underground<br />

storage organs (A. scaphoides), a vigorous creeping root<br />

system (A. cicer – chickpea milkvetch), or long taproots<br />

(A. ampullarioides – Shivwits milkvetch) (Jennifer Gremer,<br />

unpublished data; Mark Miller, personal observation;<br />

Horvath 2002). These extensive rooting systems<br />

exhibited in the genus may be linked to observed patterns<br />

of vegetative dormancy. A. ripleyi (Ripley’s milk-<br />

Figure 1. Known range of Astragalus rusbyi in northern Arizona based on herbarium collections and research studies<br />

(figure compiled by J.E. Crouse and D. L. Crisp).<br />

158


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

vetch), a species of ponderosa pine forests in northcentral<br />

New Mexico and south-central Colorado, reproduces<br />

by seed, but plants tend to allocate resources toward<br />

survival of individual plants, and it is believed to<br />

build up root stock reserves when aboveground parts are<br />

consumed (Ladyman 2003). Direct evidence supports an<br />

ability to lie dormant for two years, but monitoring has<br />

not yet been used to establish if it can remain dormant<br />

for a longer period.<br />

Although a number of Astragalus species contain<br />

toxins (such as miserotoxin or swainsonine) or accumulate<br />

selenium, thus making them poisonous to livestock,<br />

there are also many species that are highly desirable to<br />

herbivores. Lesica (1995) found fecundity losses in A.<br />

scaphoides due to livestock and insect herbivory ranging<br />

from 14-90% at two sites. Observations have confirmed<br />

that inflorescences are consumed by ants and<br />

moth larvae. Loss of seeds to weevil predation ranged<br />

from 0-33%. Sugars such as those found in flower nectar<br />

may increase palatability. Lesica also found very low<br />

recruitment, accounting for less than 17% of population<br />

growth. However, despite heavy losses in reproductive<br />

output and low recruitment, populations can continue to<br />

persist and increase in size. He suspects that persistence<br />

of many populations of long-lived plants may be more<br />

reliant on growth and survival of established plants than<br />

on recruitment from seed. Herbivory by cattle and game<br />

has also been observed in A. terminalis (railhead milkvetch),<br />

and seed predation in A. ripleyi may be the cause<br />

of significant seed loss (Heidel and Vanderhorst 1996,<br />

Ladyman 2003). Apparently, like A. scaphoides, this<br />

species has low recruitment rates and allocates a significant<br />

amount of resources toward maintenance of the<br />

root system. A. ripleyi is also consumed by a number of<br />

arthropods (aphids, treehoppers, carpenter ants), rodents<br />

and large mammals, including cattle, elk, deer, sheep<br />

and goats. A ninety percent reduction in fruit production<br />

due to herbivory was observed in A. ampullarioides<br />

(Shivwits milkvetch), which the authors suggest could<br />

have a significant impact on reproductive output (Miller<br />

et al. 2007). The toxicity of A. rusbyi is unknown.<br />

The effects of disturbances, such as tree thinning or<br />

burning, on Astragalus species vary widely. A. ripleyi is<br />

thought to be a “fire evader” rather than a stress tolerator<br />

(Ladyman 2003). Following fire, plants have been<br />

observed in areas where they have not been detected<br />

before, presumably emerging from dormant root systems<br />

underground. However, the stress-tolerator category<br />

may be appropriate, for a pattern of rapid colonization<br />

following fire and drought has also been observed<br />

in this species (Ladyman 2003).<br />

Thinning activities in pinyon-juniper woodlands at<br />

Mesa Verde National Park appeared to cause an increase<br />

in Poa fendleriana (muttongrass) that could result in<br />

undesirable competition impacts on A. schmolliae<br />

(Anderson 2004). Grass seeding post-fire also has the<br />

potential to cause negative impacts on this species.<br />

Drought is deleterious, but it is likely tolerant of fire<br />

because of a deep taproot. However, monitoring indicates<br />

that while fire may confer short-term benefits, it<br />

may also have long-term detrimental impacts (Anderson<br />

2004).<br />

OBSERVATIONS FROM FIELD STUDIES<br />

A. rusbyi has a very small range in northern Arizona,<br />

with the bulk of its population limited to a band approximately<br />

18 x 7 km (11 x 4.5 mi) in size to the west<br />

and north of the San Francisco Peaks and a few scattered<br />

populations to the west (Figure 1). Some of its<br />

habitat has been subjected to large wildfires over the last<br />

few decades; other areas have undergone ecological restoration<br />

treatments (tree thinning and prescribed burning);<br />

and much of its range in ponderosa pine forest is<br />

slated to undergo such treatments in the near future. Increasing<br />

tree densities of ponderosa pine, and a cessation<br />

of frequent fires in ponderosa pine forests since<br />

Euro-American settlement of this area of the Southwest<br />

have been well documented (Covington and Moore<br />

1994, Fulé et al. 1997).<br />

We currently do not have a thorough understanding<br />

of the basic ecology of this species. Additionally, we<br />

have insufficient knowledge of the effects of increased<br />

tree densities, tree thinning, or fire on the population<br />

dynamics. However, some limited information is available<br />

from large landscape scale studies within its range.<br />

Fisher and Fulé (2004) installed 121 20x50 m permanent<br />

monitoring plots on the south side of the San Francisco<br />

Peaks (specifically Agassiz Peak). Plots were established<br />

in five forest types: ponderosa pine, mixed<br />

conifer, aspen, spruce/fir and bristlecone pine. Overstory<br />

measurements and plant community data were<br />

collected between 2000 and 2003. A. rusbyi was found<br />

to be an indicator species for ponderosa pine forest, with<br />

an indicator value of 36.5 (p


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 2. Frequency of Astragalus rusbyi by treatment over time at G.A. Pearson Natural Area near Flagstaff, AZ.<br />

treatments (G.A. Pearson Natural Area), Laughlin and<br />

others (2008) found A. rusbyi to be an indicator species<br />

of both thinned treatments and treatments that involved<br />

thinning plus burning (mean indicator value of 25.0;<br />

p


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

D.C. Laughlin (personal communication, 2009) collected<br />

trait data on 137 ponderosa pine understory species,<br />

including A. rusbyi, and found that, on average, it<br />

has a higher specific leaf area (24 mm 2 /mg) and higher<br />

nitrogen and phosphorous concentrations in its foliar<br />

tissue (4.4% and 0.18%, respectively) (Figure 4). Because<br />

of the high nitrogen content, it has a relatively<br />

high net photosynthetic rate. On average, it also has a<br />

lower leaf dry matter content (0.19 mg/mg) and foliar<br />

C:N mass ratio (10.7). Combined with its high photosynthetic<br />

rate and comparatively tall stature (with an<br />

average height of 31 cm), it is able to compete well with<br />

other understory species, but is not very tolerant of<br />

stresses such as deep shade. The combined trait data<br />

place it in the category of a competitive ruderal species<br />

(Hodgson et al. 1999).<br />

CONCLUSIONS<br />

As previously mentioned, little is known of the ecology<br />

of this locally abundant but narrowly endemic species,<br />

and much of its known range is slated to undergo<br />

various thinning and prescribed burning activities<br />

in the very near future. Many Astragalus species are<br />

long-lived, recruit slowly by seed, and maintain longlived<br />

seeds in the soil seed bank. Whether A. rusbyi utilizes<br />

a similar strategy is unknown but could be determined<br />

from additional research. We currently do not<br />

have a thorough understanding of the population dynamics<br />

of this species over time. Rigorous long-term<br />

demographic monitoring would be valuable in determining<br />

population baselines and is essential for understanding<br />

the ecology and conservation and habitat needs of<br />

this species. Such monitoring can also reveal patterns<br />

that might be caused by precipitation fluctuations. From<br />

the information available, it appears to have a large taproot,<br />

which should give some resistance to the impacts<br />

of drought and fire, but high-intensity fire or burning at<br />

peak growth times could be detrimental. It has shown<br />

positive to no effects from tree thinning and prescribed<br />

burning operations in ecological restoration research<br />

studies, but additional research that specifically targets<br />

this species would be useful before we can draw firm<br />

conclusions.<br />

Figure 3. Proportion of permanent monitoring plots through time containing Astragalus rusbyi at an ecological restoration<br />

study area near Flagstaff, AZ. Treatments were randomly assigned within each block and included (a) no thinning,<br />

no burning (control), (b) 1.5-3 tree replacement (high-intensity thinning), (c) 2-4 tree replacement (mediumintensity<br />

thinning), and (d) 3-6 tree replacement (low-intensity thinning). All treatment units were thinned in 1999<br />

and subsequently treated with prescribed fire in spring 2000 (Block 3) and spring 2001 (Blocks 1 and 2).<br />

161


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 4. Astragalus rusbyi is located at the positive end of the ‘leaf economic spectrum’ because of its high specific<br />

leaf area and tissue nitrogen concentration. This scatterplot illustrates the results of a principal components analysis<br />

of 133 plant species and 10 functional traits (adapted from Laughlin 2009) that occur in southwestern USA ponderosa<br />

pine forests. Species symbols are coded by plant functional types. A few species names are highlighted: ARPU =<br />

Aristida purpurea, BLTR = Blepharoneuron tricholepis, CAGE = Carex geophila, HECO = Hesperostipa comata,<br />

IRMI = Iris missouriensis, LUAR = Lupinus argenteus, MUMO = Muhlenbergia montana, OXLA = Oxytropis lambertii,<br />

PIPO = Pinus ponderosa, QUGA = Quercus gambelii.<br />

ACKNOWLEDGEMENTS<br />

We would like to thank Wally Covington, Margaret<br />

Moore, Pete Fulé, Marta Fisher, Julie Korb, Mark<br />

Daniels, Jon Bakker, and Cheryl Casey and the staff and<br />

students of the Northern Arizona University Ecological<br />

Restoration Institute and School of Forestry who assisted<br />

with data collection. We would also like to thank<br />

Martin Wojciechowski and Jill Craig for their reviews<br />

of the paper and the U.S. Forest Service for providing<br />

funding to carry out the majority of the research described<br />

in this paper.<br />

REFERENCES<br />

Allphin, L., N. Brian and T. Matheson. 2005. Reproductive<br />

success and genetic divergence among varieties<br />

of the rare and endangered Astragalus cremnophylax<br />

(Fabaceae) from Arizona, USA. Conservation Genetics.<br />

6: 803-821.<br />

Anderson, D.G. 2004. Populations Status Survey of<br />

Schmoll’s Milkvetch (Astragalus schmolliae C.L. Porter).<br />

Prepared for National Park Service Mesa Verde<br />

National Park. Colorado Natural Heritage Program. Fort<br />

Collins, CO. 81 pp.<br />

Barneby, R.C. 1964. Atlas of North American Astragalus.<br />

Memoirs of The New York Botanical Garden.<br />

13:1–1188.<br />

162


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Barneby, R.C. 1989. Fabales. Intermountain Flora,<br />

vol. 3, part B. A. Cronquist et al. (editors). The New<br />

York Botanical Garden, New York.<br />

Breinholt, J.W., R.Van Buren, O.R. Kopp, and C.L.<br />

Stephen. 2009. Population genetic structure of an endangered<br />

<strong>Utah</strong> endemic, Astragalus ampullarioides (Fabaceae).<br />

American Journal of Botany. 96(3):661-667.<br />

Covington, W.W. and M.M. Moore. 1994. Southwestern<br />

ponderosa pine forest structure: changes since<br />

Euro-American settlement. Journal of Forestry. 92: 39-<br />

47.<br />

Fisher, M.A. and P.Z. Fulé. 2004. Changes in forest<br />

vegetation and arbuscular mycorrhizae along a steep<br />

elevation gradient in Arizona. Forest Ecology and Management.<br />

200: 293–311.<br />

Fulé, P.Z, W. W. Covington and M.M. Moore. 1997.<br />

Determining reference conditions for ecosystem management<br />

of southwestern ponderosa pine forests. Ecological<br />

Applications.7: 895-908.<br />

Gamon, J.G. 1995. Report on the status of Astragalus<br />

sinuatus Piper. Olympia, WA: Washington Natural<br />

Heritage Program, Washington Department of Natural<br />

Resources.<br />

Greene, H.L. 1884. New plants of the Pacific Coast.<br />

Bulletin of the California Academy of Sciences. 1(1 –<br />

Feb. 1884): 8.<br />

Gremer, J.R., E.E. Crone, and P. Lesica. <strong>2012</strong>. Are<br />

dormant plants hedging their bets? Demographic consequences<br />

of prolonged dormancy in variable environments.<br />

The American Naturalist 179(3):315-327.<br />

Heidel, B.L. and J. Vanderhorst. 1996. Sensitive<br />

plant surveys in Beaverhead and Madison counties, MT.<br />

Unpublished report to the Bureau of Land Management.<br />

Montana Natural Heritage Program, Helena. 85 pp. plus<br />

appendices.<br />

Hodgson, J.G., P.J. Wilson, R. Hunt, J.P. Grime, and<br />

K. Thompson. 1999. Allocating C-S-R plant functional<br />

types: a soft approach to a hard problem. Oikos. 85:282-<br />

294.<br />

Horvath, J. 2002. Factsheet Astragalus cicer (Cicer<br />

milkvetch). University of Saskatchewan, Rangeland<br />

Ecosystems and <strong>Plant</strong>s, PL EC 434.3.<br />

Karron, J.D., Y.B. Linhart, C.A. Chaulk, and C.A.<br />

Robertson. 1988. Genetic structure of populations of<br />

geographically restricted and widespread species of Astragalus<br />

(Fabaceae). American Journal of Botany. 75<br />

(8): 1114-1119.<br />

Ladyman, J.A.R. 2003. Astragalus ripleyi Barneby<br />

(Ripley’s milkvetch): A technical conservation assessment.<br />

Prepared for the USDA Forest Service, Rocky<br />

Mountain Region, Species Conservation Project. 48 pp.<br />

Laughlin, D.C., J.D. Bakker, M.L. Daniels, M.M.<br />

Moore, C.A. Casey, J.D. Springer. 2008. Restoring plant<br />

species diversity and community composition in a ponderosa<br />

pine-bunchgrass ecosystem. <strong>Plant</strong> Ecology.<br />

197:139-151.<br />

Lesica, P. 1995. Demography of Astragalus scaphoides<br />

and effects of herbivory on population growth.<br />

Great Basin Naturalist. 55(2): 142-150.<br />

Lesica, P. and B.M. Steele. 1994. Prolonged dormancy<br />

in vascular plants and implications for monitoring<br />

studies. Natural Areas Journal. 14(3): 209-212.<br />

Lesica, P., R. Yurkewycz, and E.E. Crone. 2006.<br />

Rare plants are common where you find them. American<br />

Journal of Botany. 93(3): 454-459.<br />

Miller, M.E., R.K. Mann and J.D. Yount. 2007. Ecological<br />

Investigations of the Federally Endangered<br />

Shivwits Milk-Vetch (Astragalus ampullarioides) –<br />

2006 Annual Report. U.S. Department of the Interior,<br />

U.S. Geological Survey, in cooperation with the National<br />

Park Service, Zion National Park. Open-File Report<br />

2007-1050. 34 pp.<br />

NatureServe. 2009. NatureServe Explorer: An online<br />

encyclopedia of life [web application]. Version 7.1.<br />

NatureServe, Arlington, Virginia. Available http://<br />

www.natureserve.org/explorer. (Accessed: May 27,<br />

2009 ).<br />

Sanderson, M.J. 1991. Phylogenetic Relationships<br />

within North American Astragalus L. (Fabaceae). Systematic<br />

Botany. 16(3): 414-430.<br />

Spellenberg, R. 1974. Chromosome number as an<br />

indication of relationships of Astragalus, section Strigulosi<br />

(Leguminosae), with descriptive notes on A. altus.<br />

The Southwestern Naturalist. 18(4): 393-396.<br />

Spellenberg, R. 1976. Chromosome numbers and<br />

their cytotaxonomic significance for North American<br />

Astragalus (Fabaceae). Taxon. 25(4): 463-476.<br />

Welsh, S.L. 2007. North American Species of Astragalus<br />

Linnaeus (Leguminosae): A Taxonomic Revision.<br />

Brigham Young University, Provo, UT. 932 pp.<br />

Wojciechowski, M.F. 2005. Astragalus (Fabaceae):<br />

A molecular phylogenetic perspective. Brittonia. 57(4):<br />

382-396.<br />

163


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Long-term Responses of Penstemon clutei (Sunset Crater Beardtongue)<br />

to Root Trenching and Prescribed Fire: Clues for Population Persistence<br />

Judith D. Springer 1 , Peter Z. Fulé 2 , and David W. Huffman 1<br />

1 Ecological Restoration Institute, Northern Arizona University, Flagstaff, AZ, and<br />

2 School of Forestry, Northern Arizona University, Flagstaff, AZ<br />

Abstract. Penstemon clutei A. Nelson (Sunset Crater beardtongue) is narrowly endemic to the cinder hills and volcanic<br />

fields northeast of Flagstaff, Arizona. Disturbances such as wildfire, tornadoes, logging activity, and tree mortality<br />

from bark beetle outbreaks appear to stimulate regeneration of this species, but the manner in which populations<br />

persist between events is still largely unknown. From 1994-2000, we examined P. clutei responses to prescribed<br />

burning and root trenching treatments that were experimentally implemented as proxies for surface fire and reduced<br />

tree densities that might be observed following natural disturbance. We revisited this experiment in 2008 to assess<br />

long-term effects of the treatments. We also collected soil samples at this time to evaluate the importance of a persistent<br />

seed bank in population dynamics. In 2008, the mean number of P. clutei plants on trenched plots had declined<br />

with time, but was still significantly higher than on the control plots (mean density of 7.4 plants in trenched plots vs.<br />

0.6 plants in control plots). There was no significant difference in density between burned and unburned plots. Only<br />

21 P. clutei seedlings emerged from 176 soil seed bank samples, and we found no correlation between the number of<br />

P. clutei plants aboveground and the number of emergents from the samples. A targeted study to obtain samples near<br />

the base of reproductively mature plants produced 9 emergents from 30 samples. Results from this work suggest that<br />

disturbances that reduce competition for soil resources may be associated with long-term population persistence. Latent<br />

seed banks appear to be of only minor importance in recovery after disturbance; however, additional research<br />

with larger sample sizes would allow for greater confidence in this conclusion. We also recommend that additional<br />

long-term research be conducted on the response of this species to specific disturbances and stressors such as wildfire,<br />

tree mortality from bark beetle outbreaks, and water limitations.<br />

Penstemon clutei (Sunset Crater beardtongue) is a<br />

narrow endemic that occurs on volcanic soils to the<br />

northeast of Flagstaff in northern Arizona. The species<br />

is primarily restricted to tephra deposits from the Sunset<br />

Crater eruption (estimated dates of eruption vary from<br />

approximately 1040-1100 AD) at an elevation of approximately<br />

2135 m (7000 ft), but a disjunct population<br />

is also present on older cinder cones about 20 km to the<br />

northwest of Sunset Crater (Figure 1). P. clutei is typically<br />

found in open ponderosa pine forests and pinyonjuniper<br />

woodlands in areas containing a sparse understory,<br />

commonly on fairly coarse and dry, cindery soils<br />

that lie over a series of finer textured sandy or silty<br />

bands, which may alternate with coarse layers of cinders<br />

(Abella and Covington 2006, Phillips et al. 1992). The<br />

type specimen was collected by Willard Clute in July<br />

1923 north of the San Francisco Peaks in “lava sand,”<br />

and was described and named by Aven Nelson from the<br />

University of Wyoming (Nelson 1927). Growing to<br />

about 50-75 cm (20-30 in) in height, P. clutei has bluish-green<br />

glaucous leaves with serrated margins and<br />

gradually inflated, deep pink corollas. It is a very showy<br />

and attractive specimen plant and is readily available to<br />

gardeners through the horticultural trade. Flowering<br />

times vary by year, but it has been observed to flower<br />

from April through early September. It is ranked G2<br />

(imperiled) by NatureServe (2009) and is on the U.S.<br />

Forest Service sensitive species list for Region 3<br />

(Southwestern Region) (Arizona Game and Fish Department<br />

2003; D. C. Crisp, personal communication,<br />

2009).<br />

There is speculation that P. clutei is descended from<br />

P. pseudospectabilis (desert penstemon) and that geographic<br />

isolation occurred following the Sunset Crater<br />

eruption (Bateman 1980), or it may be intermediate between<br />

P. pseudospectabilis and P. palmeri (Palmer’s<br />

penstemon) (Clokey and Keck 1939). Phylogeny reconstruction<br />

of the genus Penstemon using nuclear and<br />

chloroplast sequence data and parsimony analysis produced<br />

incongruent results (Wolfe et al. 2006). Strict<br />

consensus trees generated from ITS (Internal Transcribed<br />

Spacers) placed P. clutei in a polytomy with P.<br />

bicolor (pinto beardtongue), P. floridus (Panamint<br />

beardtongue), P. palmeri, and P. rubicundus (Wassuk<br />

Range beardtongue). In contrast, strict consensus trees<br />

generated from chloroplast sequence data placed P. clutei<br />

as sister to P. centranthifolius (scarlet bugler) (Wolfe<br />

et al. 2006). With both methods (ITS and chloroplast<br />

sequence), the genera within the tribe Cheloneae had<br />

high bootstrap values. However, few terminal line-<br />

164


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 1. Range of Penstemon clutei (Sunset Crater beardtongue) in northern Arizona.<br />

ages of sister taxa within the Penstemon clade had bootstrap<br />

values above 70%, which is the generally accepted<br />

value for moderate to strong support. The contradictory<br />

results are likely due to hybridization and/or rapid speciation<br />

among penstemons. Wolfe and her co-authors<br />

(1998) have documented hybridization among some<br />

Penstemon species and have also demonstrated that pollen-mediated<br />

gene flow occurs via hummingbird vectors.<br />

Many narrow endemics are found in extreme edaphic<br />

conditions, including recent volcanic soils (Lesica et al.<br />

2006). Characteristic of many of these species are high<br />

population growth rates but poor dispersal rates. Their<br />

restricted ranges may in some cases be due more to recent<br />

evolution than to ecological tolerance: it may simply<br />

be the case that some species have not yet had time<br />

to spread across the landscape and may therefore be<br />

relatively young (neoendemics) (Lesica et al. 2006).<br />

Neoendemism is common in intermountain regions of<br />

western North America, and P. clutei is likely a fairly<br />

recently evolved species.<br />

Little is currently known about the ecology of P. clutei,<br />

and information in the scientific literature is sparse.<br />

It is believed to be a short-lived perennial like many<br />

other taxa in the genus Penstemon, perhaps living five<br />

to seven years on average. No long-term population<br />

studies following individual plants have been conducted<br />

to date in its natural habitat, so estimates of its longevity<br />

are purely speculative at this time. Observations in the<br />

field have suggested a link to disturbances. Prolific<br />

growth was observed following the Burnt Fire in 1973<br />

(Goodwin 1979) and the Hochderffer Fire in 1996 (Fulé<br />

et al. 2001). It has also been observed growing in large<br />

numbers in the path left by a tornado (Crisp 1996) as<br />

well as surrounding Pinus edulis (pinyon pine) trees that<br />

were killed by drought and bark beetles in 2002-2003<br />

(J.D. Springer, personal observations, 2008 and 2009).<br />

Phillips and others (1992) noted vigorous plants and<br />

high seedling numbers in areas of past disturbance, especially<br />

from logging operations. <strong>Plant</strong>s were particularly<br />

prevalent near decaying logs and stumps. Large<br />

numbers of reproductively mature plants are also sometimes<br />

found in a ring surrounding recent Pinus ponderosa<br />

(ponderosa pine) snags (Fulé et al. 2001).<br />

Because plants had been noted to emerge in abundance<br />

following wildfire, two prescribed burning studies<br />

165


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

were established by the U.S. Forest Service, but results<br />

were inconclusive (Nagiller 1992). We initiated a study<br />

in 1992 to test the hypothesis that restoration of historic<br />

ecosystem conditions may enhance the sustainability of<br />

this species (Fulé et al. 2001). This study encompassed<br />

several components, including seed germination studies,<br />

a field seeding trial, a prescribed burning study and a<br />

trenching experiment. We initiated the prescribed burning<br />

component in 1994 to test the hypothesis that prescribed<br />

fire would increase P. clutei density by removing<br />

litter and competing vegetation. The results suggest<br />

that prescribed burning caused a significant decline in<br />

density by as much as 75%. However, density also declined<br />

in two of the three control areas (in one area also<br />

by as much as 75%). So, while prescribed burning appears<br />

to be responsible for the death of mature plants,<br />

natural population declines may also occur in the absence<br />

of disturbance.<br />

After evaluating results from the prescribed burning<br />

experiment, we investigated the possibility that vigorous<br />

responses following fires were a result of mortality of<br />

overstory trees and removal of root competition (Fulé et<br />

al. 2001). We initiated a study in 1998 to test the hypothesis<br />

that cutting root competition through trenching<br />

would increase P. clutei density. In 1999, one year following<br />

the trenching, there was a significant difference<br />

in density between trenched (mean of 104.9 plants/plot)<br />

and control plots (14.0 plants/plot), mostly in the form<br />

of seedlings. By 2000, densities had declined to an average<br />

of 30.6 vs. 1.5 plants in the trenched and control<br />

plots, respectively, mostly due to the death of seedlings.<br />

Two preliminary conclusions were drawn from the<br />

trenching study: 1) trenching had a positive effect on P.<br />

clutei reproduction, and this trend was still evident a<br />

year later, and 2) increases in P. clutei were likely due<br />

to reduced root competition with overstory trees. Although<br />

our earlier germination experiments indicated<br />

that P. clutei did not exhibit innate seed dormancy under<br />

laboratory conditions (see Fulé et al. 2001), we were<br />

puzzled over the dramatic field response to root trenching.<br />

A field seeding trial of P. clutei showed very poor<br />

rates of establishment (0.1-0.6%), with no seedlings establishing<br />

after an April seeding, and only a minimal<br />

number establishing following an October seeding (Fulé<br />

et al. 2001). Determining whether P. clutei maintains a<br />

persistent soil seed bank is crucial for conservation and<br />

management efforts. The combined results from our<br />

previous studies suggest that it does not form a persistent<br />

seed bank, that there may be dissimilarities in germination<br />

rates between plants from different habitats,<br />

and that field emergence is extremely low and/or seedling<br />

mortality is high. Collecting P. clutei seeds from<br />

additional habitats could yield new information on<br />

whether this species exhibits cyclic dormancy patterns<br />

or dormancy that differs in contrasting habitats.<br />

It remains largely unknown, then, how long populations<br />

of P. clutei plants persist on the landscape following<br />

disturbance, what mechanism this species employs<br />

to colonize an area following disturbance, or how it is<br />

able to disperse across the landscape. In an effort to gain<br />

answers to some of these questions, we revisited the<br />

study area ten years after root trenching and 13-14 years<br />

following prescribed burning. Our objectives were to<br />

assess the long-term effects of the prescribed burning<br />

and trenching treatments and to evaluate the importance<br />

of a persistent seed bank in population dynamics.<br />

METHODS<br />

Fulé and others (2001) described methods of our previous<br />

prescribed burning and root trenching studies in<br />

detail, but we will also summarize them here. The experimental<br />

studies described in this and the 2001 paper<br />

were established in 1992-1994 and conducted on Coconino<br />

National Forest lands in the vicinity of O'Leary<br />

Peak, adjacent to Sunset Crater National Monument<br />

(Figure 1). The elevation of the study area is approximately<br />

2100-2300 m (6890-7550 ft). Soils are cindery<br />

and deep, well-drained Vitrandic Ustochrepts and Typic<br />

Ustorthents (Miller et al. 1995). Weather records from<br />

Sunset Crater National Monument, 1 km south of the<br />

study area, include an annual precipitation average of<br />

42.7 cm (16.8 in) (1969-2008), with most precipitation<br />

occurring in winter and during the summer monsoon<br />

(July-September). However, annual precipitation has<br />

varied widely in recent decades from a low of 23.6 cm<br />

(9.3 in) in 1989 to 66 cm (26.0 in) in 1992. The average<br />

minimum temperature in January is -11 o C and the average<br />

maximum temperature in July is 29 o C.<br />

We established an experiment to study the effects of<br />

prescribed burning on the P. clutei community in 1994.<br />

Forty P. clutei plant-centered plots were established,<br />

each with a 2.5 m radius circle (area = 19.6 m 2 ) centered<br />

0.3 m northwest of an existing plant. P. clutei was tallied<br />

in four categories: seedling, second-year plant, mature<br />

plant, and dead. Field experience indicated that the<br />

distinctions between the living plant categories were<br />

approximate. Plots were randomly selected for burn or<br />

control treatments, and burning was conducted in September<br />

1994. Burn season effects were tested on a second<br />

experimental site immediately north of the fall<br />

burning site, and twenty randomly selected plots were<br />

burned in April 1995. The fall burn plots were re-measured<br />

in July 1995. All 80 plots, including spring and fall<br />

burns, were re-measured in August/September 1996;<br />

August 1997; August 1998; and August/September<br />

2008. Changes in P. clutei density were analyzed with<br />

repeated measures analysis of variance (Systat 8.0,<br />

SPSS Science, Chicago). Data were square-root transformed<br />

to meet ANOVA assumptions. In 2008, data<br />

166


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

were cube-root transformed, and the software used was<br />

JMP 8.0 (SAS, Cary, NC).<br />

In October 1997 we established ten new experimental<br />

plots for root trenching within the low-elevation<br />

burning study area. Each plot was paired with a nearby<br />

control plot from the original burn experiment. The<br />

number of trenched plots was eventually dropped to<br />

eight due to off-road vehicle damage and other factors.<br />

The below ground effects of tree removal were simulated<br />

by digging a narrow trench approximately 1 m<br />

deep around each plot. Trenches were located 50 cm<br />

outside the plot boundary to avoid physical disturbance<br />

within the measured area and were lined with plastic<br />

sheeting to minimize tree root regrowth (Milne 1979).<br />

Trenches were backfilled immediately after lining, and<br />

plots were re-measured in August 1998, September<br />

1999, August 2000 and August/September 2008. P. clutei<br />

density was analyzed as described for the prescribed<br />

burning study above.<br />

In 2008, we collected two soil seed bank samples<br />

from each of the 80 prescribed burning study plots and<br />

also from the trenched plots (176 total). Samples were<br />

collected to a depth of 5 cm, approximately 15 cm away<br />

from the original plot center to the east and west, and<br />

each core was approximately 70 cm 3 in volume. We also<br />

collected 30 targeted seed bank samples 15 cm to the<br />

east and west of reproductively mature individuals that<br />

were located outside of plots. Seeds are dispersed in the<br />

fall and winter, and germination is thought to occur during<br />

late spring and early summer rains, so we collected<br />

soil seed bank samples in late August and early September,<br />

presumably after germination occurred, but before<br />

new seeds were dispersed, in an effort to capture seeds<br />

in the persistent soil seed bank. We sieved samples to<br />

remove large cinders and placed the soil samples on potting<br />

soil in gallon-sized pots. Samples were placed in<br />

the greenhouse in September 2008 and received artificial<br />

light, one application of Miracle Gro® and daily<br />

watering for five months, using the seed emergence<br />

method (Ter Heedt et al. 1996).<br />

RESULTS<br />

In 2008, thirteen and fourteen years after spring and<br />

fall burning, respectively, there was no significant difference<br />

in P. clutei density between burned and unburned<br />

plots (Figure 2). Mean density in burned plots<br />

(combined spring and fall burns) was 0.9 live plants,<br />

and mean density in control plots was 0.6 live plants.<br />

Over the course of this study, there has been a general<br />

decline in P. clutei in both burned and control plots.<br />

However, there was still a significant difference in mean<br />

density between trenched and control plots ten years<br />

Figure 2. Mean density of Penstemon clutei plants following a prescribed burn study near Sunset Crater National<br />

Monument in northern Arizona. Pre-treatment data were collected in 1994 and 1995. Bars indicate standard error.<br />

167


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 3. Mean density of Penstemon clutei plants following a root trenching experiment near Sunset Crater National<br />

Monument in northern Arizona . Pre-treatment data were collected in 1997. Bars indicate standard error (each plot<br />

was 19.6 m 2 ; n=80; p value in 2008 was


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

point to a number of studies showing that root trenching<br />

can lead to an increase in the availability of mineral nitrogen.<br />

Research by Selmants and Hart (2008) indicates<br />

that there are large carbon and nitrogen pools and fluxes<br />

under the canopies of one-seed juniper (Juniperus<br />

monosperma) in soils around Sunset Crater. Soils under<br />

juniper and pinyon pine canopies are generally higher in<br />

both carbon and nitrogen than are intercanopy sites, but<br />

these levels vary according to the age of the soils, which<br />

are of volcanic origin in this region.<br />

Abella and Covington (2006) obtained samples from<br />

a number of soil types across the Coconino National<br />

Forest, including black and red cinder soils in the vicinity<br />

of Sunset Crater, and determined that black cinder<br />

soils contain the driest surface soils among those tested.<br />

These soils are very sandy (>90% concentration at 0-15<br />

cm depth), and contained the fewest plant species per<br />

plot. Red cinder soils are also quite sandy (averaging<br />

63% concentration at 0-15 cm depth). Soil samples<br />

taken from red cinders contained no calcium carbonate,<br />

but these soils had higher organic carbon and total nitrogen<br />

than the black cinder soils, and they also had higher<br />

soil moisture. P. clutei populations in pinyon-juniper<br />

woodlands are typically found on these older, red cinder<br />

soils.<br />

The soils on which P. clutei grows, then, are arguably<br />

some of the harshest in northern Arizona and are<br />

susceptible to extreme environmental fluctuations.<br />

Seedling mortality is high, but once plants reach maturity,<br />

they have adapted to the harsh, arid environment by<br />

means of a large taproot or thick lateral roots (D.W.<br />

Huffman, personal observations, 2008) and thick leaves.<br />

The species also must have developed adaptations to be<br />

able to rapidly colonize following disturbance, perhaps<br />

through longevity, rapid dispersal, high germinability,<br />

or a persistent soil seed bank. Soil seed banks buffer<br />

populations against environmental variation, and seed<br />

dormancy is a mechanism of escape from unfavorable<br />

conditions in time (compared to dispersal, which is an<br />

escape in space) (Doak et al. 2002). Short-lived perennials<br />

in areas of high environmental variation, which includes<br />

most rare plants in the Southwest, often rely on<br />

the soil seed bank for recruitment (Doak et al. 2002). P.<br />

lemhiensis (Lemhi penstemon) and P. palmeri have<br />

been documented to buffer populations against environmental<br />

fluctuations by maintaining a persistent soil seed<br />

bank (Heidel and Shelly 2001, Meyer and Kitchen<br />

1992). Conversely, long-lived perennials are often more<br />

reliant on growth and survival of established plants than<br />

on recruitment from seed or soil seed banks (Lesica<br />

1995). If a species does not exhibit innate dormancy, it<br />

is unlikely that it forms a soil seed bank. Because P.<br />

clutei plants have been observed to appear in large numbers<br />

following a disturbance such as the Hochderffer<br />

Fire (P.Z. Fulé, personal observation, 1997 and 1998),<br />

conventional thinking is that this species forms a persistent<br />

soil seed bank (Phillips et al. 1992), but it may also<br />

maintain genetic diversity through existing reproductively<br />

mature plants scattered across the landscape, or<br />

exhibit rapid dispersal rates following disturbance.<br />

While it does seem from our study that P. clutei may<br />

form a minor persistent seed bank, the degree of its importance<br />

in recovery following disturbance is unknown,<br />

and larger sample sizes from additional habitats are necessary<br />

in order to make inferences about its significance<br />

for recruitment following disturbance.<br />

Meyer et al. (1995) found a diversity of germination<br />

timing mechanisms in 38 Intermountain West Penstemon<br />

species. Seeds of many of these species diverge<br />

into two fractions. One fraction does not exhibit dormancy<br />

and will germinate readily under optimal conditions<br />

in the first year. The other fraction may respond to<br />

chilling cues and become nongerminable, allowing for<br />

between-year carryover in the soil seed bank. Meyer and<br />

her co-authors (1995) found this strategy to be especially<br />

common in populations of penstemons from middle<br />

elevation areas that have unpredictable winters.<br />

Meyer and Kitchen (1992) discovered that P. palmeri<br />

seeds undergo cyclic dormancy changes in the field.<br />

Moist chilling induces secondary dormancy in about<br />

half of the seeds, while moisture combined with summer<br />

temperatures removes secondary dormancy. These<br />

mechanisms allow for a persistent soil seed bank and for<br />

seeds that can persist from year to year without burial.<br />

The result is that some seeds germinate in the spring,<br />

while those seeds that are rendered dormant by chilling<br />

are carried over in the soil seed bank. Another fruitful<br />

area of research for this species could include seed augmentation<br />

studies to determine if a paucity of viable<br />

seeds may be limiting establishment. Abella (2008) conducted<br />

such a study with P. virgatus (upright blue<br />

beardtongue) in a ponderosa pine forest not far from our<br />

study site and found that under particular experimental<br />

conditions, the site environment (e.g., tree overstory)<br />

apparently was more limiting to recruitment than either<br />

leaf litter thickness or seed availability.<br />

Our results also indicate that prescribed burning<br />

alone does not seem to be a useful management tool for<br />

this species, as it appears to kill reproductively mature<br />

individuals leading to potential decreases in available<br />

seeds for future recruitment. An experiment involving<br />

use of prescribed fire as a management tool for P. lemhiensis<br />

returned variable results (Heidel and Shelly<br />

2001). Fire appeared to cause mortality of adult plants<br />

ranging from 25-75%. However, the burning caused<br />

an increased recruitment rate of 4600-6400%. As we<br />

pointed out in our previous paper (Fulé et al. 2001),<br />

patchy tree mortality does appear to benefit P. clutei.<br />

Tree mortality from the 2002-2003 bark beetle outbreak<br />

among pinyon pines appears to be correlated with dra-<br />

169


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

matic increases in the number of reproductively mature<br />

individuals on the south-facing slopes of cinder cones in<br />

the vicinity of Sunset Crater and Indian Flat (J.D.<br />

Springer, personal observations, 2008 and 2009).<br />

Although some P. clutei populations contain hundreds<br />

or thousands of individuals, populations are often<br />

widely dispersed, and there are a few major threats that<br />

could jeopardize this species in the future. The entire<br />

range of P. clutei has not yet been mapped; however, a<br />

large portion of its known range falls within the Cinder<br />

Hills OHV (off-highway vehicle) area (Figure 1). Most<br />

of this area is not fenced and OHV use spills outside the<br />

boundaries shown in the map. Although P. clutei appears<br />

to benefit from disturbance, whether disturbance<br />

is beneficial or detrimental depends on the type of disturbance<br />

and the amount of impact. No quantifiable data<br />

has yet been collected on the impacts of OHV activity<br />

on this species, but anecdotal evidence points to OHVs<br />

as a direct factor in adult P. clutei mortality (J.D.<br />

Springer, personal observations, 2008). OHV activity<br />

causes above- and belowground soil impacts, resulting<br />

in decreased soil moisture, increased soil bulk density,<br />

and increased water infiltration time, which have been<br />

shown to negatively impact plant species in the area,<br />

such as ponderosa pine (Kennedy 2005).<br />

While OHV use and impacts can be controlled, potential<br />

negative changes to P. clutei habitat from climate<br />

change cannot. Climate models predict a more arid climate<br />

in the southwestern U.S. in the coming decades<br />

(Seager et al. 2007). This species already lives in a<br />

harsh environment, and any major decreases in available<br />

soil moisture could significantly impact its long-term<br />

viability. Additional threats include potential hybridization<br />

with other Penstemon species brought to the area<br />

for horticulture or highway revegetation purposes, herbivory,<br />

insect damage and urban expansion.<br />

Determining the long-term population dynamics of<br />

this species is integral to future conservation management<br />

planning and points out the direct need for longterm<br />

monitoring, particularly in the face of potential<br />

climate change and unmanaged OHV use in the center<br />

of its habitat. Teasing out whether P. clutei population<br />

declines occur from disturbance, absence of disturbance,<br />

senescence, competition, drought, climate change, interactive<br />

effects, or other as yet undetermined factors will<br />

be critical for understanding future conservation and<br />

management needs for this species.<br />

ACKNOWLDEGMENTS<br />

We thank Deb Crisp, Barb Phillips, and Frank Thomas<br />

(Coconino National Forest), Steve Rosenstock<br />

(Arizona Game and Fish Department), Paul Whitefield<br />

(National Park Service), and Susie Smith (Northern Arizona<br />

University) for their assistance in gathering locations<br />

and for their input on ecology and experimental<br />

design; Scott Abella (UNLV and Public Lands Institute)<br />

and Nancy Brian (National Park Service) for their review<br />

of the manuscript; and Joe Crouse, Mark Daniels,<br />

Chris McGlone, Mike Stoddard, Chelsea Green, Cat<br />

McGowan, Don Normandin and students and staff of<br />

the Ecological Restoration Institute at Northern Arizona<br />

University for assistance with data collection, mapping,<br />

greenhouse study maintenance and statistical analysis.<br />

LITERATURE CITED<br />

Abella, S.R. 2008. <strong>Plant</strong> recruitment in a northern<br />

Arizona ponderosa pine forest: testing seed- and leaf<br />

litter-limitation hypotheses. Pp. 119-127 in Olberding,<br />

Susan D., and Moore, Margaret M., tech coords. 2008.<br />

Fort Valley Experimental Forest—A Century of Research<br />

1908-2008. Proceedings RMRS-P-53CD. Fort<br />

Collins, CO: U.S. Department of Agriculture, Forest<br />

Service, Rocky Mountain Research Station. 408 p.<br />

Abella, S.R. and W.W. Covington. 2006. Forest ecosystems<br />

of an Arizona Pinus ponderosa landscape: multifactor<br />

classification and implications for ecological<br />

restoration. Journal of Biogeography. 33: 1368-1383.<br />

Arizona Game and Fish Department. 2003. Penstemon<br />

clutei. Unpublished abstract compiled and edited<br />

by the Heritage Data Management System, Arizona<br />

Game and Fish Department, Phoenix, AZ. 5 pp.<br />

Bateman, Gary C. 1980. Natural Resource Survey and<br />

Analysis of Sunset Crater and Wupatki National Monuments.<br />

Final Report (Phase III). Prepared for the Office<br />

of Natural Resources Management, Southwest Region,<br />

National Park Service.<br />

Clokey, I.W. and D.D. Keck. 1939. Reconsideration<br />

of certain members of Penstemon subsection spectabilis.<br />

Bulletin of the Southern California Academy of Science.<br />

38: 8-13.<br />

Coomes, D.A. and P.J. Grubb. 2000. Impacts of Root<br />

Competition in Forests and Woodlands: A Theoretical<br />

Framework and Review of Experiments. Ecological<br />

Monographs. 70(2): 171-207.<br />

Crisp, D.L. 1996. Monitoring of Penstemon clutei A.<br />

Nels. on tornado salvage. In USDA Forest Service General<br />

Technical Report RM-GTR-283, pp. 243-246.<br />

Rocky Mountain Forest and Range Experiment Station,<br />

Fort Collins, Colorado.<br />

Doak, D.F., D. Thomson, and E.S. Jules. 2002. Population<br />

Viability Analysis for <strong>Plant</strong>s: Understanding the<br />

Demographic Consequence of Seed Banks for Population<br />

Health. In: Steven R. Beissinger and Dale R.<br />

McCullough, Eds. Population Viability Analysis. University<br />

of Chicago Press.<br />

Fulé, P.Z., J.D. Springer, D.W. Huffman, and W.W.<br />

Covington. 2001. Response of a rare endemic, Penstemon<br />

clutei, to burning and reduced belowground competition.<br />

Pp 139-152 in Maschinski. J., and L. Holter<br />

170


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

(tech. coords.), Proceedings: Third Rare and Endangered<br />

<strong>Plant</strong> Conference. Proceedings RMRS-P-23.<br />

Goodwin, G. 1979. Observations on Penstemon clutei<br />

on the Coconino National Forest. Unpublished report<br />

on file at Supervisor's Office, Coconino National Forest,<br />

Flagstaff, AZ. 7 pp.<br />

Heidel, B. and J.S. Shelly. 2001. The Effects of Fire<br />

on Lemhi Penstemon (Penstemon lemhiensis) – Final<br />

Monitoring Report 1995-2000. Montana Natural Heritage<br />

Program. Prepared for Beaverhead-Deerlodge National<br />

Forest and Dillon Field Office – Bureau of Land<br />

Management.<br />

Kennedy, K.J. 2005. Above- and belowground impacts<br />

of off-road vehicles negatively affect establishment<br />

of a dominant forest tree. Master’s thesis, Northern<br />

Arizona University, Flagstaff. 39 pp.<br />

Lesica, P. 1995. Demography of Astragalus scaphoides<br />

and effects of herbivory on population growth.<br />

Great Basin Naturalist. 55(2): 142-150.<br />

Lesica, P., R. Yurkewycz, and E.E. Crone. 2006.<br />

Rare plants are common where you find them. American<br />

Journal of Botany. 93(3): 454-459.<br />

Meyer, S.E., and S.G. Kitchen. 1992. Cyclic seed<br />

dormancy in the short-lived perennial Penstemon<br />

palmeri. Journal of Ecology 80: 115-122.<br />

Meyer, S.E., S.G. Kitchen and S.L. Carlson. 1995.<br />

Seed germination timing patterns in intermountain Penstemon<br />

(Scrophulariaceae). American Journal of Botany.<br />

82(3): 377-389.<br />

Miller, G., N. Ambos, P. Boness, D. Reyher, G.<br />

Robertson, K. Scalzone, R. Steinke and T. Subirge.<br />

1995. Terrestrial ecosystems survey of the Coconino<br />

National Forest. USDA Forest Service, Southwestern<br />

Region.<br />

Milne, M.M. 1979. The effects of burning, root<br />

trenching, and shading on mineral soil nutrients in<br />

southwestern ponderosa pine. Master’s thesis, Northern<br />

Arizona University, Flagstaff.<br />

Nagiller, S. 1992. Untitled report of prescribed burning<br />

results on file at Elden Ranger District, Coconino<br />

National Forest, Flagstaff, AZ. 2 pp.<br />

NatureServe. 2009. NatureServe Explorer: An online<br />

encyclopedia of life [web application]. Version 7.1.<br />

NatureServe, Arlington, Virginia. Available http://<br />

www.natureserve.org/explorer. (Accessed: May 27,<br />

2009 ).<br />

Nelson, A. 1927. A new Penstemon from Arizona.<br />

American Botanist. 33: 109-110.<br />

Phillips, A. M. III, M. B. Murov, and R. J. van Ommeren.<br />

1992. Distribution and ecology of Sunset Crater<br />

beard tongue (Penstemon clutei) in the Cinder Hills<br />

area, Coconino National Forest, Arizona. Unpublished<br />

report on file at the Supervisor's Office, Coconino National<br />

Forest, Flagstaff, Arizona. 11 pp.<br />

Seager, R., M. Ting, I. Held, Y. Kushnir, J. Lu, G.<br />

Vecchi, H. Huang, N. Harnik, A. Leetmaa, N. Lau, C.<br />

Li, J. Velez, and N. Naik. 2007. Model Projections of an<br />

Imminent Transition to a More Arid Climate in Southwestern<br />

North America. Science. 316(5828): 1181-<br />

1184.<br />

Selmants, P.C. and S.C. Hart. 2008. Substrate age<br />

and tree islands influence carbon and nitrogen dynamics<br />

across a retrogressive semiarid chronosequence. Global<br />

Biogeochemical Cycles. 22, GB1021, doi:<br />

10.1029/2007GB003062, 2008.<br />

Ter Heedt, G.N.J., G.L. Verwiej, R.M. Bekker and<br />

J.P. Bakker. 1996. An improved method for seed-bank<br />

analysis: seedling emergence after removing the soil by<br />

sieving. Functional Ecology. 10(1): 144-151.<br />

Wolfe, A.D., Q. Ziang and S.R. Kephart. 1998. Assessing<br />

hybridization in natural populations of Penstemon<br />

(Scrophulariaceae) using hypervariable intersimple<br />

sequence repeat (ISSR) bands. Molecular Ecology.<br />

7:1107-1125.<br />

Wolfe, A.D., C.P. Randle, S.L. Datwyler, J.J.<br />

Morawetz, N. Arguedas and J. Diaz. 2006. Phylogeny,<br />

taxonomic affinities and biogeography of Penstemon<br />

(<strong>Plant</strong>aginaceae) based on ITS and cpDNA sequence<br />

data. American Journal of Botany. 93(11): 1699-1713.<br />

171


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

¡Viva thamnophila! Ecology of Zapata Bladderpod<br />

(Physaria thamnophila), an Endangered <strong>Plant</strong> of the<br />

Texas-Mexico Borderlands<br />

Dana M. Price,<br />

US Army Corps of Engineers, Albuquerque District<br />

Christopher F. Best,<br />

US Fish and Wildlife Service, Austin Ecological Services Field Office<br />

Norma L. Fowler,<br />

Section of Integrative Biology, University of Texas at Austin,<br />

and Alice L. Hempel,<br />

Department of Biology and Health Sciences, Texas A&M University- Kingsville<br />

Abstract. Conserving rare plants is dependent upon our ability to identify, manage and restore their habitat. We examined<br />

the plant community associates and habitat requirements of Physaria thamnophila, an endangered herbaceous<br />

perennial of the Tamaulipan shrubland of south Texas, at four sites from 2003 to 2007. At each site, vegetation<br />

height structure and species composition were sampled concurrently with intermittent censuses of P. thamnophila.<br />

We found significant and interesting differences among sites and years, as well as between our quantitative results<br />

and previous descriptions of P. thamnophila’s community and habitat. Existing plant community descriptions and<br />

mapped soil types do not provide a close match with our field observations. Finally, we discuss the application of<br />

these results to conservation of P. thamnophila and restoration of its community.<br />

An endangered plant that is a member of a fragmented<br />

and altered remnant plant community poses<br />

many challenges to conservation workers. In addition to<br />

threats to its (usually) few and small known populations,<br />

two additional challenges are faced. One, the detection<br />

of additional populations is often hampered by uncertainty<br />

in identifying its habitat; and two, uncertainty<br />

about its habitat requirements hampers management and<br />

restoration efforts. Physaria thamnophila (Zapata bladderpod;<br />

Brassicaceae) is one such plant. This endangered<br />

herbaceous perennial of south Texas grows in<br />

remnants of Tamaulipan thornscrub, likely on a specific<br />

but as yet poorly defined soil type and geologic substrate.<br />

Prior to the present study, the structure and composition<br />

of the specific plant community in which Zapata<br />

bladderpod occurs had been examined quantitatively<br />

for only one site (Sternberg 2005).<br />

The U.S. Fish and Wildlife Service (USFWS) Recovery<br />

Plan (USFWS 2004) highlights the need for better<br />

knowledge of the habitat and community in which P.<br />

thamnophila lives for several purposes, including discovering<br />

additional populations, locating sites for establishing<br />

new populations, and restoring and managing its<br />

habitat. This study addresses these goals by providing a<br />

detailed, quantitative description of the habitat and<br />

community of four sites with persistent (as defined by<br />

NatureServe 2002) P. thamnophila populations. The<br />

results will not only be useful for P. thamnophila conservation,<br />

but can improve our ability to manage and<br />

restore one of the communities that comprise the<br />

Tamaulipan thornscrub ecosystem.<br />

SPECIES<br />

Physaria thamnophila (Rollins and E.A. Shaw)<br />

O’Kane and Al-Shehbaz (formerly Lesquerella thamnophila)<br />

is a federally listed endangered species (USFWS<br />

2004) with a global conservation ranking of G1<br />

(critically imperiled; NatureServe 2009). It is a shortlived<br />

perennial with a rosette of silvery- or gray-green<br />

leaves covered with stellate trichomes (Figure 1) and<br />

one to several flexuous, sprawling to ascending flowering<br />

stems. The yellow flowers give rise to subglobose<br />

silicles on recurved pedicels (Rollins and Shaw 1973)<br />

(Figure 2). It usually flowers in spring (February to<br />

April), but can flower during midsummer or as late as<br />

September in response to rain (Poole et al. 2007; Sternberg<br />

2005). Germination has not been observed, but<br />

probably occurs in response to rainfall in cool weather.<br />

Under prolonged dry conditions, all of the plant’s leaves<br />

may die, making some surviving plants very difficult to<br />

locate.<br />

172


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Figure 1. Physaria thamnophila rosette leaves.<br />

SITES<br />

All known populations of P. thamnophila (eight verifiably<br />

extant, one not accessible, and one historic population;<br />

data on file, Natural Diversity Database, Texas<br />

Parks and Wildlife Department, Austin, TX) are in Starr<br />

and Zapata Counties, Texas, between Zapata and Roma,<br />

a distance of 69 km (Figure 3). All populations occur in<br />

remnant Tamaulipan thornscrub vegetation (Poole et al.<br />

2007). For this study, we examined all sites with at least<br />

100 plants to which we had access. Other sites were excluded<br />

due to small populations, lack of access to private<br />

land, or discovery too late to include in the study.<br />

Three of the four sites in this study (Arroyo Ramirez,<br />

Arroyo Morteros, and Cuellar) were in the Lower Rio<br />

Grande Valley National Wildlife Refuge. The fourth<br />

study site, Santa Margarita, was on private land. All<br />

four sites were within 18 km of each other in Starr<br />

County, Texas, between Falcon Dam (26° 33’ N, 99°<br />

09’W) and Roma (26° 24’N, 99° 01’W).<br />

The climate in this region is hot and often dry. At<br />

Falcon Dam, the nearest station in the National Oceanic<br />

and Atmospheric Administration (NOAA) data base, the<br />

months of July and August have the highest means of<br />

daily maximum temperatures (37.6 o C and 37.5 o C, respectively)<br />

and January has the lowest mean daily minimum<br />

temperature (7.7 o C) (NOAA 2009). Between<br />

1963 and 2007, average annual precipitation at Falcon<br />

Dam was 515 mm. However, annual precipitation varied<br />

widely, as during those 45 years, there were 7 years<br />

with more than 600 mm and 9 years with less than 400<br />

mm annual precipitation. During our study, 2003 and<br />

2004 were wet (804 mm and 671 mm, respectively) and<br />

2005 was dry (353 mm). However, precipitation in the<br />

area occurs largely as isolated thunderstorms, and likely<br />

varied among sites.<br />

Most known populations occur on the edges of terraces<br />

above the Rio Grande flood plain. Physaria thamnophila<br />

populations have been found on the Jackson,<br />

173<br />

Figure 2. Physaria thamnophila fruiting stem.<br />

Yegua and Laredo formations (Bureau of Economic Geology<br />

1976; Poole et al. 2007; USFWS 2004), all of<br />

which are Eocene calcareous sandstones and clays. All<br />

four study sites were on Eocene sandstones: Cuellar and<br />

Arroyo Morteros on the Yegua Formation, and Arroyo<br />

Ramirez and Santa Margarita on the Jackson Group<br />

(Bureau of Economic Geology 1976). Arroyo Morteros,<br />

Arroyo Ramirez, and Santa Margarita were located<br />

along the edge of the cliff that marks the edge of the<br />

flood plain of the Rio Grande; Cuellar was ~ 3000 m<br />

from the flood plain in an area without a cliff.<br />

Known P. thamnophila populations occur on shallow,<br />

well-drained sandy loam soil. Soils at known sites<br />

are mapped as members of the Zapata, Maverick, Catarina,<br />

or Copita series as described by the Natural Resources<br />

Conservation Service (NRCS) (Poole et al.<br />

2007; USFWS 2004). These highly calcareous soils are<br />

derived from Eocene sandstone, clay and shale. Catarina<br />

soils are derived from Frio and Yegua formation parent<br />

material; these and Maverick soils contain up to 15 percent<br />

gypsum. Copita soil is derived from weakly consolidated<br />

calcareous sandstone of the Jackson Formation<br />

and is only slightly (2 percent) gypsiferous (NRCS<br />

2009; Thompson et al. 1972). The soils at our study<br />

sites have been described as follows: Arroyo Morteros:<br />

Copita, Zapata and Catarina; Arroyo Ramirez: Jimenez-


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 3: Physaria thamnophila distribution, Starr and Zapata Counties, Texas.<br />

174


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Quemado; Cuellar: Catarina; Santa Margarita: Maverick,<br />

eroded and Jimenez-Quemado (NRCS 2009). However,<br />

soils maps for these areas lack precision and the<br />

inclusion of P. thamnophila sites within Jimenez-<br />

Quemado soil polygons is incorrect (see Discussion). In<br />

some sites, fossil oyster shells or gypsum crystals were<br />

conspicuous.<br />

METHODS<br />

Vegetation Data Collection<br />

Vegetation data were collected once at Arroyo Ramirez,<br />

Arroyo Morteros, and Santa Margarita and twice at<br />

Cuellar in conjunction with intermittent censuses of the<br />

study sites (Fowler et al. 2011). Cuellar vegetation was<br />

sampled in 2002 and 2007, Arroyo Ramirez in 2003,<br />

Santa Margarita in 2004, and Arroyo Morteros in 2007.<br />

Due to limitations of time and personnel, we set up<br />

study plots at one site each year in 2002-2005. In each<br />

site, permanent circular plots marked with rebar were<br />

located randomly along transects. Some of the Cuellar<br />

census plots did not have vegetation data collected in<br />

one or both years and were dropped from analyses.<br />

At Cuellar, 30 plots were located in an uncleared<br />

area and 30 plots in an area that had been brush-cleared<br />

using a ‘Woodgator©’ (roller-chopper) in <strong>December</strong><br />

2000. This decreased shrub canopy (see Results) and<br />

significantly increased herbaceous species richness and<br />

grass abundance in the cleared area (Fowler et al. 2011).<br />

A transect ran along the margin between the two areas,<br />

with the plots on either side. In each of the other three<br />

sites, transects were located to sample the entire Physaria<br />

thamnophila population. The full extent of the<br />

Santa Margarita population was unknown at the beginning<br />

of this study, and the Arroyo Ramirez and Arroyo<br />

Morteros populations were discovered in fall 2002 and<br />

summer 2004, respectively. Therefore, we first conducted<br />

reconnaissance surveys to map the populations’<br />

extents using GPS. The two roughly linear populations<br />

that followed rocky outcrops, Arroyo Ramirez and<br />

Santa Margarita, were each sampled by running a discontinuous<br />

transect (excluding unoccupied habitat)<br />

along the long axis, through the approximate midline of<br />

the population. The transect length was then divided<br />

into 30 strata, with plots located within each stratum at<br />

random distances either side of the transect line. There<br />

were 30 plots at Santa Margarita (Figure 4) and 34 plots<br />

at Arroyo Ramirez. At Arroyo Morteros, where the<br />

population occupied a large, irregular polygon, we<br />

placed 58 plots along ten parallel transects of different<br />

lengths, spaced 38m apart. The initial plot in each transect<br />

was located a random distance between 0.1 m to 10<br />

m along the transect, with successive plots spaced 13.8<br />

m apart.<br />

By 2007, the cleared portion of Cuellar was quite<br />

similar to the other three sites in herbaceous species<br />

richness and grass abundance (Fowler et al. 2011). Arroyo<br />

Morteros had on average somewhat greater herbaceous<br />

species richness than the other sites (12 species<br />

per plot, on average, versus 6 to 10 in the other sites;<br />

Fowler et al. 2011).<br />

Vegetation data from each plot were collected in five<br />

circular subplots. Each subplot was 0.25m in radius.<br />

One of these subplots was centered on the plot’s central<br />

point, and the other four were centered around points<br />

located on the circumference of the circle at the distal<br />

ends of four radial lines emanating from the central<br />

point (two parallel and two perpendicular to the transect).<br />

In each subplot, the presence/absence of each species<br />

was recorded in each of 4 height categories. These<br />

categories were 0.0 to 0.5 m, 0.5 to 1.0 m, 1.0 to 2.0 m,<br />

and 2.0 to 3.0 m above ground. No plants were taller<br />

than 3 m. Species names follow USDA PLANTS database<br />

(<strong>2012</strong>) with the exception of Physaria and Paysonia,<br />

for which we follow the treatment of Al-Shehbaz<br />

and O’Kane (2002). Most common species, particularly<br />

woody dominants, were identified in the field. We collected<br />

voucher specimens when species were encountered<br />

in flower; specimens are being prepared for deposit<br />

at the University of Texas herbarium (TEX-LL).<br />

It was not always possible to definitively identify<br />

species at the time when they were first observed. As a<br />

result, in a few cases the interim identifications used for<br />

unknown species were not consistent between years.<br />

Therefore, for analysis we pooled (a) the two Chamaesyce<br />

(Euphorbiaceae) species; (b) Chaetopappa bellioides<br />

and Aphanostephus skirrhobasis (Asteraceae);<br />

and (c) Chamaesaracha sordida and Physalis cinerascens<br />

(Solanaceae). This reduced the number of 'species'<br />

in the analyses from 150 to 147. For simplicity, each of<br />

these three pairs of species is referred to simply as a<br />

species. ‘Bare ground’ was recorded as a 148 th ‘species’.<br />

A few plants could never be identified; almost all of<br />

these were without fruits or flowers and many were<br />

grasses. These have been left as unknown species 1,<br />

unknown grass, etc. in Table 1.<br />

The number of subplots in each plot in which the<br />

species occurred was used to quantify the abundance of<br />

each species. The abundance of a given species in a plot<br />

therefore had a value between 0 and 5. For P. thamnophila<br />

only, counts of numbers of individuals in each plot<br />

were also available. However, censuses of P. thamnophila<br />

were conducted in all sites in 2006 and 2007 only.<br />

We calculated the density of P. thamnophila in each<br />

plot by dividing the number of individuals in the plot by<br />

the plot’s area, for each plot separately. P. thamnophila<br />

densities in 2006 and in 2007 were then averaged, for<br />

each census plot separately.<br />

175


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 4. Sampling design for discontinuous sites.<br />

176


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. Species composition in Physaria thamnophila study plots by site<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Abutilon fruticosum Malvaceae 1 - - 0.74 - - - -<br />

Acacia berlandieri Fabaceae 1 - - - - - - 0.67<br />

Acacia rigidula Fabaceae X 5 47.20 33.06 60.00 39.26 46.55 24.12 18.00<br />

Acalypha monostachya Euphorbiaceae 2 - - - - - 1.18 9.33<br />

Acleisanthes obtusa Nyctaginaceae 2 - - 2.22 - - 2.35 -<br />

Acourtia runcinata Asteraceae 1 - - - - - - 0.67<br />

Allionia incarnata Nyctaginaceae 3 - - - - 0.34 8.82 3.33<br />

Aloysia macrostachya Verbenaceae X 3 - 2.42 - 13.33 - 0.59 3.33<br />

Argythamnia humilis var.<br />

humilis<br />

Euphorbiaceae 5 0.80 - 0.74 2.22 0.69 1.18 0.67<br />

Aristida purpurea Poaceae 5 17.60 1.61 28.89 4.44 11.72 42.94 66.67<br />

Aristolochia sp. Aristolochiaceae 1 - - - - - - 0.67<br />

Astragalus nuttallianus Fabaceae 2 - - 0.74 2.22 - - -<br />

Astragalus sp. 2 Fabaceae 1 - - - - 0.34 - -<br />

Ayenia pilosa Sterculiaceae 5 - - 2.96 2.22 0.34 4.12 2.67<br />

Bahia absinthifolia Asteraceae 1 - - - - 0.69 - -<br />

Bothriochloa laguroides<br />

ssp. torreyana<br />

Poaceae 1 - - - - - 1.18 -<br />

Bouteloua repens Poaceae 1 - - - - - - 0.67<br />

Bouteloua trifida Poaceae 5 16.80 - 8.15 1.48 9.66 8.24 2.67<br />

Cardiospermum dissectum<br />

Sapindaceae 1 - - - - - 16.47 -<br />

Celtis ehrenbergiana Ulmaceae X 2 - - 0.74 - 1.03 - -<br />

Celtis laevigata Ulmaceae 1 - - - - - 0.59 -<br />

Cenchrus spinifex Poaceae 2 - - 2.22 - 0.69 - -<br />

Cevallia sinuata Loasaceae 1 - - - - 3.45 - -<br />

Chaetopappa bellioides<br />

and Aphanostephus skirrhobasis<br />

var. kidderi<br />

Chamaesaracha sordida<br />

and Physalis cinerascens<br />

Asteraceae 5 1.60 - 19.26 3.70 50.00 4.12 14.67<br />

Solanaceae 3 - - - 4.44 26.55 - 8.67<br />

177


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. continued<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Chamaesyce laredana and<br />

C. cinerascens<br />

Euphorbiaceae 4 24.00 14.52 5.93 11.85 6.55 - 6.00<br />

Cissus trifoliata Vitaceae 2 - - - 0.74 0.69 - -<br />

Citharexylum brachyanthum<br />

Verbenaceae X 4 0.80 0.81 1.48 2.96 2.07 0.59 -<br />

Commelina erecta Commelinaceae 4 - - - 1.48 0.69 0.59 2.67<br />

Condalia spathulata Rhamnaceae 1 - - - - 0.34 - -<br />

Convolvulus equitans Convolvulaceae 3 - - 0.74 0.74 0.69 - -<br />

Cooperia sp. Liliaceae 2 - - - - - 1.76 1.33<br />

Croton incanus Euphorbiaceae X 1 - - - - - 12.94 -<br />

Cynanchum barbigerum Asclepiadaceae X 4 - - 1.48 2.96 1.72 2.94 -<br />

Cyperus sp. Cyperaceae 5 0.80 - 2.96 0.74 6.21 1.18 1.33<br />

Dalea nana Fabaceae 2 4.00 - 8.89 - - 0.59 -<br />

Dalea pogonathera Fabaceae 1 - - - - 0.34 - -<br />

Digitaria cognata Poaceae 4 - - 2.96 - 0.34 7.65 16.00<br />

Diospyros texana Ebenaceae X 4 0.80 - 0.74 - 3.45 2.35 4.00<br />

Echinocactus texensis Cactaceae 1 - - - - - 0.59 -<br />

Echinocereus enneacanthus<br />

Cactaceae 3 - - - - 0.34 1.76 0.67<br />

Echinocereus poselgeri Cactaceae 4 - 0.81 - 1.48 0.34 1.76 0.67<br />

Echinocereus reichenbachii<br />

ssp. fitchii<br />

Cactaceae 1 - - - - - - 0.67<br />

Ephedra antisyphilitica Ephedraceae X 3 0.80 - 0.74 - - 18.24 4.00<br />

Eragrostis curtipedicellata<br />

Poaeae 3 - - 0.74 0.74 3.79 - -<br />

Eriogonum greggii Polygonaceae 1 - - - - - - 26.67<br />

Erioneuron pilosum Poaceae 2 - - - - - 4.12 0.67<br />

Escobaria emskoetteriana Cactaceae 2 - - - - - 0.59 0.67<br />

Evolvulus alsinoides Convolvulaceae 3 - - 8.15 9.63 8.97 - -<br />

Eysenhardtia texana Fabaceae X 5 - - 2.22 1.48 6.21 7.65 4.67<br />

178


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. continued<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Ferocactus hamatacanthus<br />

var. sinuatus<br />

Cactaceae 2 - - - - - 0.59 2.00<br />

Fleischmannia incarnata Asteraceae 1 - - - - 0.34 - -<br />

Florestina tripteris Asteraceae 1 - - - - 0.34 - -<br />

Forestiera angustifolia Oleaceae X 5 0.80 8.06 2.22 8.15 23.10 4.71 6.00<br />

Galium sp. Rubiaceae 4 - - 5.93 4.44 1.38 - 9.33<br />

Galphimia angustifolia Malphigiaceae 4 - - 2.22 1.48 3.79 - 2.00<br />

Gamochaeta pensylvanica Asteraceae 1 - - - - 1.03 - -<br />

Gilia incisa Polemoniaceae 1 - - 2.22 - - - -<br />

Grusonia schottii Cactaceae 1 0.80 - - - - - -<br />

Guaiacum angustifolium Zygophyllaceae 4 3.20 2.42 2.96 1.48 - 3.53 1.33<br />

Heliotropium confertifolium<br />

Heliotropium curassavicum<br />

Boraginaceae 2 - - - - 1.03 - 3.33<br />

Boraginaceae 1 - - - - 0.34 - -<br />

Herissantia crispa Malvaceae 2 - - - - 0.69 1.18 -<br />

Heterotheca sp. Asteraceae 1 - - - - 0.34 - -<br />

Hibiscus martianus Malvaceae 3 - - 0.74 0.74 - 0.59 -<br />

Ibervillea lindheimeri Cucurbitaceae 2 - - - - 0.34 - 0.67<br />

Jatropha dioica Eupohorbiaceae X 5 4.80 3.23 5.93 7.41 11.38 5.88 3.33<br />

Jefea brevifolia Asteraceae 2 0.80 - - - - - 1.33<br />

Justicia pilosella Acanthaceae 2 - - - - 0.34 1.18 -<br />

Karwinskia humboldtiana Rhamnaceae X 5 1.60 5.65 4.44 9.63 7.93 12.35 16.00<br />

Koeberlinia spinosa Capparaceae X 1 - 0.81 - - - - -<br />

Krameria ramosissima Krameriaceae 5 8.00 4.84 7.41 7.41 3.45 16.47 10.00<br />

Lantana achyranthifolia Verbenaceae 1 - - - - 6.90 - -<br />

Lantana urticoides Verbenaceae 2 - - - 1.48 1.72 - -<br />

Lepidium lasiocarpum<br />

var. wrightii<br />

Brassicaceae 1 - - - - 2.76 - -<br />

Leucophyllum frutescens Scrophulariaceae X 5 30.40 52.42 62.96 72.59 13.79 11.76 11.33<br />

179


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. continued<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Linum lundellii Linaceae 5 - - 14.07 10.37 2.07 1.18 2.67<br />

Lippia graveolens Verbenaceae X 5 6.40 12.10 13.33 19.26 3.10 22.94 8.00<br />

Lupinus texensis Fabaceae 1 - - - - - - 1.33<br />

Lycium berlandieri Solanaceae 2 - - - - 0.34 0.59 -<br />

Macrosiphonia lanuginosa<br />

Apocynaceae 1 - - - - - - 0.67<br />

Mammillaria heyderi Cactaceae 1 - - - - - 2.35 -<br />

Mammillaria sphaerica Cactaceae 1 - - - - - 0.59 -<br />

Manfreda longiflora Agavaceae 1 - - - - - 0.59 -<br />

Maurandya antirrhiniflora<br />

Scrophulariaceae X 1 - - - - 6.55 - -<br />

Melampodium cinereum Asteraceae 5 25.60 9.68 29.63 27.41 23.79 3.53 4.67<br />

Melinis repens Poaceae 1 - - - - 0.34 - -<br />

Mimosa texana<br />

(M. wherryana)<br />

Fabaceae X 5 8.80 13.71 13.33 13.33 11.72 5.29 12.67<br />

Nama hispidum Hydrophyllaceae 4 - - 20.74 28.89 41.38 - 1.33<br />

Oenothera laciniata Onagraceae 3 - - 8.89 7.41 10.34 - -<br />

Opuntia engelmannii Cactaceae X 4 0.80 - 1.48 0.74 2.41 2.94 -<br />

Opuntia leptocaulis Cactaceae X 4 - 1.61 - 1.48 0.34 1.18 0.67<br />

Opuntia sp. Cactaceae 1 - - - - - 0.59 -<br />

Oxalis dichondrifolia Oxalidaceae 3 - - 5.19 2.22 1.38 - -<br />

Palafoxia texana Asteraceae 2 - - 1.48 - 7.24 - -<br />

Panicum hallii Poaceae 1 - - - - 0.69 - -<br />

Parietaria pensylvanica Urticaceae 4 - - 3.70 1.48 9.66 - 1.33<br />

Parkinsonia texana var.<br />

texana (Cercidium texanum)<br />

Fabaceae X 4 1.60 3.23 4.44 2.96 3.10 0.59 -<br />

Parthenium confertum Asteraceae 3 4.00 2.42 2.96 3.70 22.76 - -<br />

Paspalum setaceum Poaceae 1 - - - - 0.34 - -<br />

Passiflora tenuiloba Passifloraceae 2 - - - - 0.34 2.35 -<br />

180


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. continued<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Paysonia lasiocarpa Brassicaceae 1 - - - - 1.72 -<br />

Pennisetum ciliare Poaceae 3 - - 0.74 - 11.38 9.41 -<br />

Phacelia congesta Hydrophyllaceae 1 - - - - 2.07 - -<br />

Phemeranthus aurantiacus<br />

Portulacaceae 2 - - - 0.74 0.34 - -<br />

Phyllanthus polygonoides Euphorbiaceae 5 - - 21.48 9.63 8.97 7.06 4.67<br />

Physaria thamnophila Brassicaceae 5 20.80 1.61 22.22 5.93 24.14 25.29 17.33<br />

<strong>Plant</strong>ago hookeriana <strong>Plant</strong>aginaceae 3 - 0.81 2.96 2.22 9.66 - -<br />

Polygala lindheimeri Polygalaceae 5 30.40 50.00 24.44 48.15 5.52 20.59 14.67<br />

Portulaca sp. Portulaceae 1 - - - - 1.72 - -<br />

Rivina humilis Phytolaceae 1 - - - - - 1.76 -<br />

Salvia ballotiflora Lamiaceae 1 - - - - 0.69 - -<br />

Schaefferia cuneifolia Celastraceae X 3 2.40 0.81 2.22 - 0.34 - -<br />

Senna bauhinioides Fabaceae 3 - - 1.48 2.96 0.69 - -<br />

Setaria leucopila Poaceae 4 0.80 - - - 1.03 1.76 0.67<br />

Setaria ramiseta Poaceae 2 - - 6.67 4.44 - - -<br />

Setaria texana Poaceae 1 - - - - 7.93 - -<br />

Sida abutifolia Malvaceae 3 - - 7.41 6.67 4.48 - -<br />

Sideroxylon celastrinum Sapotaceae X 3 - - - - 0.34 11.76 4.67<br />

Sonchus oleraceus Asteraceae 2 - - 1.48 0.74 - - -<br />

Spermolepis echinata Apiaceae 1 - - - - 2.07 - -<br />

Sporobolus cryptandrus Poaceae 5 6.40 - 5.19 1.48 1.72 5.88 3.33<br />

Synthlipsis greggii Brassicaceae 3 - - - - 4.14 2.94 10.00<br />

Tetraclea coulteri Verbenaceae 3 - - 2.96 0.74 2.41 - -<br />

Thamnosma texana Rutaceae 5 28.80 11.29 39.26 28.15 6.90 10.00 7.33<br />

Thymophylla pentachaeta Asteraceae 5 38.40 3.23 71.11 54.07 50.00 11.18 5.33<br />

Tiquilia canescens Boraginaceae 4 21.60 6.45 13.33 11.85 6.55 0.59 -<br />

Tridens muticus Poaceae 5 18.40 1.61 7.41 0.74 2.07 12.94 14.67<br />

Turnera diffusa Turneraceae 1 - - - - - 58.82 -<br />

181


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. continued<br />

Scientific Name Family >1<br />

m<br />

tall?<br />

#<br />

Sites<br />

w/sp a<br />

CL02<br />

cl<br />

% of Subplots Occupied in Each Site b<br />

CL02<br />

un<br />

CL07<br />

cl<br />

CL07<br />

un<br />

AM07 AR03 SM04<br />

Unidentified cactus seedling<br />

Cactaceae 1 - - - - - - 0.67<br />

Unknown grass Poaceae 5 8.00 0.81 25.93 24.44 21.03 11.18 1.33<br />

Unknown legume Fabaceae 1 - - - 0.74 - - -<br />

Unknown 1 1 1.60 - - - - - -<br />

Unknown 2 2 3.20 2.42 0.74 1.48 - - -<br />

Unknown 3 1 1.60 - - - - - -<br />

Unknown 4 1 0.80 - - - - - -<br />

Unknown 5 5 2.40 1.61 4.44 7.41 0.69 5.29 0.67<br />

Unknown 6 1 - - - - - - 1.33<br />

Unknown 7 1 - - - - - 1.76 -<br />

Urochloa ciliatissima Poaceae 1 - - - - 1.03 - -<br />

Urochloa texana Poaceae 1 - - 0.74 - - - -<br />

Verbena sp 1. Verbenaceae 2 - - 2.22 - 2.76 - -<br />

Verbena sp 2. Verbenaceae 3 - - 2.22 0.74 1.72 - -<br />

Wedelia texana (A. Gray)<br />

B.L. Turner<br />

Asteraceae 5 2.40 - 19.26 0.74 4.48 0.59 0.67<br />

Yucca treculeana Agavaceae X 2 - - - 0.74 0.34 - -<br />

Zanthoxylum fagara Rutaceae X 1 - - - - 1.03 - -<br />

Ziziphus obtusifolia Rhamnaceae 4 - - 0.74 - 0.69 0.59 0.67<br />

Bare ground (no plants in<br />

subplot)<br />

5 0.80 7.26 - - 0.34 4.12 3.33<br />

a<br />

<strong>Number</strong> of sites with this species: Cuellar cleared and uncleared counted as different ‘sites’ due to<br />

treatment difference<br />

b Site codes: CL cl = Cuellar cleared; CL un = Cuellar uncleared; AM = Arroyo Morteros; AR = Arroyo<br />

Ramirez; SM = Santa Margarita. <strong>Number</strong>s 02-07 refer to year of vegetation data collection.<br />

182


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

RESULTS<br />

Differences in Vegetation Among Sites<br />

A total of 150 vascular plant species occurred in at<br />

least one vegetation plot in a least one site (Table 1).<br />

Sixty-six of these species occurred in only one of the<br />

four sites (pooling Cuellar treatments). While the vegetation<br />

of the four sites was similar in many ways, the<br />

relative abundances of species differed among sites<br />

(Table 1). In Arroyo Morteros, Thymophylla pentachaeta,<br />

Acacia rigidula, and Chaetopappa bellioides/<br />

Aphanostephus skirrhobasis var. kidderi were most<br />

abundant; in Arroyo Ramirez, Aristida purpurea,<br />

Turnera diffusa, Acacia rigidula, and Polygala lindheimeri;<br />

in Santa Margarita, Aristida purpurea and<br />

Eriogonum greggii; and in Cuellar, Acacia rigidula,<br />

Leucophyllum frutescens, Polygala lindheimeri, and<br />

Melampodium cinereum. Sites also differed in the abundances<br />

of less common species (Table 1). A MANOVA<br />

(multivariate analysis of variance) comparing the abundances<br />

of all species (except P. thamnophila and ‘bare<br />

ground’, for a total of 146 ‘species’ after the pooling<br />

described in the Methods) among the five site x treatment<br />

combinations (Cuellar treatments not pooled) was<br />

highly significant (Cuellar 2007 vegetation data used;<br />

Hotelling-Lawley Trace, F = 9.44, df = 548, 114.71,<br />

P < 0.0001).<br />

The four sites also differed in vegetation structure<br />

(Figure 5). The lack of vegetation above 1m in the<br />

cleared portion of Cuellar in 2002 was due to its treatment,<br />

while the uncleared portion of Cuellar had the<br />

densest canopies. The regrowth in the Cuellar cleared<br />

portion is apparent in a comparison of the 2002 and<br />

2007 graphs (Figure 5). Among uncleared sites, Arroyo<br />

Ramirez had the most open canopies, and Arroyo<br />

Morteros the tallest.<br />

Relationships Between P. thamnophila Presence and<br />

Neighboring Species<br />

Twenty-nine species, including P. thamnophila, were<br />

present in 25 percent or more of the subplots. These species,<br />

except for P. thamnophila, were used as the dependent<br />

variables in a MANOVA. In this MANOVA,<br />

site and P. thamnophila presence/absence in the 2007<br />

census were the independent variables. Cuellar 2007<br />

vegetation data were used in this analysis and the two<br />

Cuellar treatments were considered to be 2 different<br />

‘sites’, for a total of 5 site-treatment combinations. Both<br />

of the independent variables were highly significant<br />

(site-treatment combination: Hotelling-Lawley Trace F<br />

= 20.67; df = 112,476; P < 0.0001; presence/absence:<br />

Hotelling-Lawley Trace F = 1.93; df = 28,143; P =<br />

0.0066).<br />

Because the MANOVA found significant effects of<br />

P. thamnophila presence we followed it with ANOVAs<br />

(univariate analyses of variance). Six of 28 ANOVAs on<br />

the same 28 species had P-values less than 0.05 associated<br />

with the effect of P. thamnophila presence/absence.<br />

In each of these ANOVAs, presence/absence, df = 1,<br />

was the second factor in a hierarchical sums of squares<br />

table that had site, df = 3, as the first factor. Thus these<br />

P-values reflect the amount of variation that presence/<br />

absence accounted for above-and-beyond the variation<br />

accounted for by site-year combination. Plots that had<br />

P. thamnophila had on average more Diospyros texana<br />

(P = 0.0424), Acacia rigidula (P = 0.0043), unidentified<br />

seedling grasses (P = 0.0068), Mimosa texana (P =<br />

0.0012), and Chamaesaracha sordida + Physalis cinerascens<br />

(P = 0.0293), and less Tiquilia canescens (P =<br />

0.0293), than plots without P. thamnophila. These positive<br />

and negative associations with P. thamnophila<br />

should be regarded as tentative, due to multiple testing<br />

issues, deviations from the multivariate normal assumption,<br />

and correlations among abundances of the different<br />

species. Only Mimosa texana meets the Bonferroni criterion<br />

for significance (for 28 tests, P < 0.00183 to give<br />

an overall P < 0.05).<br />

Relationships Between P. thamnophila Density and<br />

Neighboring Species<br />

A stepwise regression was used (SAS PROC REG)<br />

with the average 2006/2007 density of P. thamnophila<br />

as the dependent variable. Site was included as a class<br />

variable (df = 3). All 148 ‘species’ except bare ground<br />

and P. thamnophila were available to the regression procedure.<br />

A criterion of P < 0.05 was used to retain variables<br />

in the model. Eighteen taxa met this criterion.<br />

Thirteen of them were negatively associated with P.<br />

thamnophila: Acleisanthes obtusa, Acourtia runcinata,<br />

Citharexylum brachyanthum, Thymophylla pentachaeta,<br />

Echinocactus texensis, Echinocereus reichenbachii ssp.<br />

fitchii, Galphimia angustifolia, Nama hispidum, Oenothera<br />

laciniata, Portulaca sp., Sideroxylon celastrinum,<br />

Synthlipsis greggii, and Tridens muticus. Five were<br />

positively associated with P. thamnophila: Acacia rigidula,<br />

Cyperus sp., Evolvulus alsinoides, Melampodium<br />

cinereum, and Passiflora tenuiloba.<br />

DISCUSSION<br />

Community Composition: General<br />

Each of the study sites had a rich plant community<br />

that contained shrubs, forbs, graminoids, and cacti<br />

(Table 1). The shrubs Acacia rigidula and Leucophyllum<br />

frutescens dominated all four sites. While all four<br />

study sites were similar enough to be described as having<br />

the same plant community, there were some interesting<br />

differences among them. These include Turnera diffusa,<br />

found in 59 percent of samples at Arroyo Ramirez<br />

but absent from other sites; Eriogonum greggii, itself a<br />

183


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 5: Canopy height structure at Physaria thamnophila<br />

sites.<br />

184


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

somewhat rare plant, found in 27 percent of subplots at<br />

Santa Margarita only; and the shrub Forestiera angustifolia,<br />

which occurred in 23 percent of subplots at Arroyo<br />

Morteros, but was infrequent at other sites. Since<br />

there was no evidence that P. thamnophila was declining<br />

at any of the sites, each of these slightly different<br />

plant communities is evidently suitable for P. thamnophila.<br />

Community Composition: Comparisons with Other<br />

Studies<br />

The plant community of our study sites, quantified in<br />

Table 1, does not match previously identified plant communities<br />

of the region. It also differs in some important<br />

details from previously published lists of the plant species<br />

associated with Physaria thamnophila. These differences<br />

may be attributed to site selection, scale of observation,<br />

method of community description (sampling<br />

vs. observation of visual dominants), and, for herbaceous<br />

species, year effects.<br />

The Recovery Plan (USFWS 2004) stated that P.<br />

thamnophila occurs in “an open Leucophyllum frutescens<br />

(cenizo) - Acacia berlandieri (guajillo) shrubland<br />

alliance” (Acacia rigidula - Leucophyllum frutescens -<br />

Acacia berlandieri Shrubland Alliance; NatureServe<br />

2009). The plant community we observed may be a<br />

component association of this alliance. However, A.<br />

berlanderi was rarely encountered in our study. Among<br />

the plant associations described by NatureServe (2009),<br />

the best match to the community we observed is the<br />

Acacia rigidula - Leucophyllum frutescens - Acacia berlandieri<br />

Shrubland Association (CEGL007759). However,<br />

this “broadly defined type” is not described in sufficient<br />

detail to make it useful in defining P. thamnophila<br />

habitat. Our sites somewhat resembled the Acacia<br />

rigidula - Leucophyllum frutescens - Hechtia glomerata<br />

Shrubland Association (CEGL007760; NatureServe<br />

2009). However, this association occurs on saline clay<br />

soils. Furthermore, Hechtia glomerata, while observed<br />

in the vicinity of P. thamnophila sites, was not encountered<br />

in study plots. The more general Acacia rigidula<br />

Shrubland Association (CEGL003874) is too broadly<br />

defined to be useful in identifying P. thamnophila habitat.<br />

Jahrsdoerfer and Leslie (1988) described a “Chihuahuan<br />

(or Falcon) thorn forest” community in this<br />

area, but listed only two dominant species: A. rigidula<br />

and Mimosa biuncifera. M. biuncifera does not occur in<br />

this area, but the name has been misapplied to M. texana;<br />

and thus, it is likely they were referring to M. texana.<br />

Other plant community classifications for South<br />

Texas, including McLendon (1991) and Bezanson<br />

(2000), describe associations similar to NatureServe’s.<br />

These communities include species that were not important<br />

in our study, and occur on different substrates.<br />

NRCS Ecological Site Descriptions also list different<br />

species mixes than those we encountered. Copita soils<br />

belong to the Gray Sandy Loam Ecological Site; Catarina,<br />

to the Saline Clay Ecological Site, and Mavericks<br />

to the Rolling Hardland Ecological Site. The plant communities<br />

of these ecological sites are described as having<br />

more mid-grasses and different forbs than our study<br />

sites, as well as a different mix of dominant shrubs.<br />

Jimenez and Quemado soils belong to the undescribed<br />

Gravelly Ridge Ecological Site, while their associated<br />

“unnamed, minor components” have not been assigned<br />

to an ecological site. Zapata soils are in the undescribed<br />

Shallow Ridge Ecological Site (NRCS 2009)<br />

Wu and Smeins (1999) listed sixteen woody species<br />

associated with P. thamnophila, including A. rigidula,<br />

L. frutescens, Karwinskia humboldtiana, Krameria<br />

ramosissima, and Jatropha dioica, all of which were<br />

frequently encountered in our plots. However, they also<br />

listed species that we did not find to be closely associated<br />

with Zapata bladderpod, including Prosopis glandulosa,<br />

Hechtia glomerata, Acacia berlandieri, and A.<br />

greggii. They noted a diverse herbaceous understory but<br />

did not list species. Although their report does not describe<br />

the method for characterizing vegetation, we suspect<br />

that this study reported visual dominants or species<br />

common in the general area, but not necessarily associated<br />

with P. thamnophila at a finer scale.<br />

Sternberg (2005) listed 22 woody and herbaceous<br />

species associated with P. thamnophila at Cuellar. Even<br />

though his sampling method covered the entire population<br />

area and ours focused on the cleared area and adjacent<br />

uncleared habitat, most of the frequently encountered<br />

species were important in both studies. However,<br />

there were some notable differences. For example,<br />

Sternberg did not report the woody species Mimosa texana,<br />

nor did he encounter several herbaceous species<br />

that we observed in 10 percent or more of subplots.<br />

Among the important herbaceous species that Sternberg<br />

did not observe, a few were most abundant in the<br />

cleared portion of Cuellar (Tridens muticus, Aristida<br />

purpurea), while others were only identified in 2007,<br />

following a relatively wet winter (for example, Phyllanthus<br />

polygonoides and the annuals Nama hispidum and<br />

Linum lundellii).<br />

Relationships Between P. thamnophila and Individual<br />

<strong>Plant</strong> Species<br />

A number of species that were common on upland<br />

soils in the region were absent or uncommon in our<br />

study sites. Acacia berlandieri, Prosopis glandulosa,<br />

Ziziphus obtusifolia, Zanthoxylum fagara, Cordia boissieri,<br />

and Lycium berlandieri are common members of<br />

upland plant communities in the region (Bezanson 2000;<br />

NatureServe 2009; NRCS 2009; USFWS 2004), but<br />

were absent or uncommon in our study sites. A. ber-<br />

185


landeri, cited as a community dominant in the Recovery<br />

Plan (USFWS 2004), appeared in the vegetation data in<br />

only one of our four study sites. This species is a dominant<br />

at Santa Margarita immediately upslope of the P.<br />

thamnophila population that is on a different soil type.<br />

Prosopis glandulosa never occurred within our study<br />

plots. The inclusion of Prosopis in the Recovery Plan<br />

was based on a site that we did not study, an abandoned<br />

trailer park which has subsequently been overtaken by<br />

invasive buffelgrass (Pennisetum ciliare), leaving only a<br />

small remnant on highway right-of-way (J.M. Poole,<br />

Texas Parks and Wildlife Department, personal communication<br />

20 July 2009). Other species mentioned in the<br />

Recovery Plan that we encountered infrequently include<br />

Celtis ehrenbergiana, Yucca treculeana, Ziziphus obtusifolia,<br />

and Guaiacum angustifolium, which are common<br />

visual dominants in the region that were not closely<br />

associated with P. thamnophila at our sites. These species<br />

may be better suited to sites with deeper soils or<br />

more favorable soil chemistry.<br />

Two common invasive grasses of the area, buffelgrass<br />

(Pennisetum ciliare (L.) Link) and Kleberg bluestem<br />

(Dichanthium annulatum (Forssk.) Stapf), are also<br />

conspicuous by their absences or low abundances in the<br />

four study sites. It seems likely that P. thamnophila, like<br />

other native grasses and forbs of south Texas (Sands et<br />

al. 2009), is out-competed by these invasive grasses.<br />

The plots in which P. ciliare was encountered should be<br />

monitored to determine its effect on P. thamnophila.<br />

Associated species’ negative or positive correlation<br />

with P. thamnophila may have a number of explanations.<br />

Some of the negative correlations may reflect very<br />

localized competition. For example, a plot in which<br />

Thymophylla pentachaeta or Nama hispidum were very<br />

abundant (especially in 2007) may have been colonized<br />

by these species in response to winter rain, and they in<br />

turn excluded P. thamnophila. Some negative correlations<br />

may reflect microsites not suitable for P. thamnophila,<br />

such as hardpan, where Tiquilia canescens was<br />

relatively common. Synthlipsis greggii seems to use different<br />

microsites and topographic positions than P.<br />

thamnophila; however, we did not quantify this observation.<br />

Acacia rigidula’s positive correlation with P. thamnophila<br />

may indicate facilitation, probably because as a<br />

legume it may create a soil patch with relatively high<br />

nitrogen content. Alternatively, its presence may indicate<br />

that the plot is not bare hardpan, but rather is favorable<br />

to vegetation in general. The other shrub that was<br />

positively correlated with P. thamnophila was the legume<br />

Mimosa texana. This shrub, although not rare, has<br />

a restricted range and may be more indicative of P.<br />

thamnophila habitat. A “characteristic species of the<br />

arid, sandy-soil Falcon Woodlands, which cover a small<br />

upland part of Starr and Zapata Counties” (Ideker 1999),<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

186<br />

M. texana was present at all study sites in 5 to14 percent<br />

of subplots.<br />

The positive correlation of P. thamnophila with perennial<br />

herbaceous species such as Melampodium<br />

cinereum and Evolvulus alsinoides is probably related to<br />

its similar microhabitat requirements and response to<br />

precipitation and shrubs. Other species with significant<br />

correlations were present in low frequency and are inconclusive.<br />

Edaphic Requirements of P. thamnophila<br />

Although P. thamnophila populations are mapped on<br />

several soil series, our observations in the field indicate<br />

very similar soils and geologic substrates at all sites. All<br />

four populations of P. thamnophila occur on a sandstone<br />

substrate (Figure 6), on yellowish, highly erodible,<br />

highly calcareous soils. All other Texas populations to<br />

which we and other observers have had access have<br />

similar yellowish sandy soils and occur on sandstone.<br />

Wu and Smeins (1999) report Copita and Zapata sandy<br />

loam soils as the substrate for P. thamnophila, and clarified<br />

that sites mapped as Catarina soils (saline, gypsiferous<br />

clay) are actually on sandy inclusions. Their analyses<br />

of soil from four P. thamnophila sites found very<br />

high calcium, high sulfur, and very low nitrogen levels.<br />

We believe that use of NRCS digital soil maps at a<br />

level of detail beyond which they were intended has led<br />

to confusion about the soils on which P. thamnophila<br />

occurs. P. thamnophila has never been found on<br />

Jimenez-Quemado soils (contra Poole 1989 and<br />

USFWS 2004). The parent material of these soils is<br />

gravelly alluvium, deposited by ancient, high-velocity<br />

streams on the high terraces over the Rio Grande<br />

(Thompson et al. 1972). The Jimenez-Quemado soil<br />

polygons contain inclusions of “unnamed, minor components”<br />

and rock outcrops. These outcrops, rather than<br />

Jimenez or Quemado soils, are likely habitat for P.<br />

thamnophila.<br />

Figure 6: Sandstone substrate with Physaria thamnophila<br />

plant.


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

The underlying geology at the sites is complex. The<br />

Yegua and Jackson Formations were deposited in part<br />

of the Gulf Coast geosyncline known as the Rio Grande<br />

Embayment during Eocene cycles of sedimentation.<br />

Land subsidence and marine transgressions and regressions<br />

produced fluctuating sea levels and deposition of<br />

“complexly interbedded sands, silts, and clays” as well<br />

as marine shales (Preston 2009). Dumble (1902) provided<br />

a detailed site-specific description of this complex<br />

stratigraphy. Reporting on outcrops of “buff sandstone”<br />

near Roma (in the vicinity of our study sites), he describes<br />

sections of “greenish-yellow clays” with gypsum,<br />

oyster beds, buff clays, sandy clays, and indurated<br />

sandstone.<br />

At two of our study sites that were on bluffs along<br />

the Rio Grande, two layers of buff to yellowish sandstone<br />

were interspersed with other substrates or oyster<br />

shell deposits. Fossil oyster shell beds occurred upslope<br />

from all four study sites. Gypsum crystals occurred on<br />

the soil surface in many places. The alternating layers of<br />

sandstone with fossil oyster shell, shale and clay may<br />

explain the presence of gypsum at the sites, even though<br />

the sandy soils (Copita and Zapata) are only weakly<br />

gypsiferous. The layers of different substrates may create<br />

microsites where water is more available due to<br />

seepage from relatively permeable layers located over<br />

impermeable or less permeable layers. Several species<br />

we encountered in this study are members of genera that<br />

are documented as tolerating gypsum, particularly<br />

Tiquilia, but also Nama, Eriogonum, and Acleisanthes<br />

(Moore and Jansen 2007). We have no evidence that P.<br />

thamnophila is a gypsum endemic; however, it tolerates<br />

gypsum.<br />

All four sites were undergoing active gully erosion<br />

to the degree that some slopes could be called badlands.<br />

All four sites also appeared to have high rates of sheet<br />

erosion as well, especially in the bare areas between<br />

shrubs. While erosion probably does not directly benefit<br />

P. thamnophila, high erosion rates likely reduce the<br />

number of competing species that can live in the site,<br />

and perhaps their densities. High erosion rates may be<br />

one factor contributing to the positive association between<br />

shrubs and P. thamnophila, especially P. thamnophila<br />

seedlings within our sites (Fowler et al. 2011). By<br />

slowing the rate of erosion in their immediate vicinity,<br />

shrubs may increase seedling survival there. High erosion<br />

rates may also explain why roller-chopping part of<br />

Cuellar increased the density of P. thamnophila there<br />

(Fowler et al. 2011), as the woody debris left by this<br />

treatment may also have reduced the erosion rate. However,<br />

we cannot exclude other positive effects that<br />

shrubs may have upon P. thamnophila.<br />

These edaphic features (high erosion rates, highly<br />

calcareous soils, perhaps the presence of gypsum) could<br />

be used to search for sites where additional populations<br />

might occur. They also provide some guidance for identifying<br />

sites suitable for introduction or reintroduction.<br />

Although we do not believe that P. thamnophila requires<br />

high soil erosion rates, the presence of gypsum,<br />

or even calcareous soils, all of these probably reduce the<br />

number and density of competing species. Ecologically,<br />

P. thamnophila is apparently a stress-tolerator (sensu<br />

Grime 1977, 2001), rather than a strong competitor or a<br />

ruderal species. This may also be true of many of its<br />

associates.<br />

Conservation Applications<br />

The Recovery Plan (USFWS 2004) called for (a)<br />

identification of sites where additional populations<br />

might occur; (b) identification of sites most suitable for<br />

attempts to establish new populations; (c) identification<br />

of tracts most appropriate for mitigation purposes,<br />

should that be necessary; (d) development of management<br />

plans; and (e) development of habitat restoration<br />

objectives. The vegetation structure and composition<br />

data presented here, especially that of Table 1 and Figure<br />

5, provide quantitative objectives for each of these<br />

tasks. Whereas visually dominant species mentioned in<br />

previous work can still be used to search for general<br />

areas of suitable native thornscrub vegetation, the species<br />

composition presented here provides a finer scale<br />

focus for identification of suitable habitat. The qualitative<br />

description of likely soil types provided above<br />

could be used to guide searches for new populations and<br />

identify sites for introduction or reintroduction. It would<br />

also be helpful to know the gypsum content of the soils.<br />

Management and restoration projects can use Table 1<br />

and Figure 5 to help set quantitative objectives.<br />

It should be noted that this study was not designed to<br />

compare sites with and without P. thamnophila, so results<br />

do not definitively identify what it is about these<br />

four sites that made them different from similar sites in<br />

the region. Additionally, we did not attempt to characterize<br />

small, remnant, disturbed sites, which are also<br />

important to the species’ conservation. In any case, with<br />

so few P. thamnophila populations, the absence of P.<br />

thamnophila from any particular site is hard to interpret.<br />

P. thamnophila may be absent from suitable sites due to<br />

chance, poor dispersal, disease, herbivory, or other factors,<br />

which is often a problem in studying endangered<br />

species (Hanski and Ovaskainen 2002). Experienced<br />

biologists will no doubt make their own judgments from<br />

the data we provide.<br />

ACKNOWLEDGEMENTS<br />

We thank the Lower Rio Grande Wildlife Refuge<br />

staff for their assistance; landowner Jorge Gonzales and<br />

family for access to Santa Margarita Ranch; and Tom<br />

and Elena Patterson, Thomas Adams, Robyn Cobb,<br />

Loretta Pressley, Jim Manhart, and Alan Pepper for<br />

187


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

assistance with field work. Cullen Hanks reviewed the<br />

status of Physaria populations and the map, and Alan<br />

Pepper, Karen Clary, Jackie Poole, and Martin Terry<br />

reviewed the manuscript. Their comments have greatly<br />

improved this work.<br />

LITERATURE CITED<br />

Al-Shehbaz, I.A. and S.L. O’Kane Jr. 2002.<br />

Lesquerella is united with Physaria (Brassicaceae). Novon<br />

12:319-329.<br />

Bezanson, D. 2000. Natural vegetation types of<br />

Texas and their representation in conservation areas.<br />

M.A. thesis, University of Texas, Austin. [Online].<br />

Available: http://abisw.org/bezanson/ [June 1, 2009].<br />

Bureau of Economic Geology. 1976. Geologic atlas<br />

of Texas, McAllen-Brownsville sheet. University of<br />

Texas, Austin, TX.<br />

Dumble, E.T. 1902. Geology of southwestern Texas.<br />

Transactions of the American Institute of Mining Engineers<br />

33:913-987.<br />

Fowler, N.L., C.F. Best, D.M. Price, and A.L. Hempel.<br />

2011. Ecological requirements of the Zapata bladderpod<br />

Physaria thamnophila, an endangered<br />

Tamaulipan thornscrub plant. Southwestern Naturalist<br />

56(3):341-352.<br />

Grime, J. P. 1977. Evidence for the existence of three<br />

primary strategies in plants and its relevance to ecological<br />

and evolutionary theory. American Naturalist<br />

111:1169-1194.<br />

Grime, J. P. 2001. <strong>Plant</strong> strategies, vegetation processes,<br />

and ecosystem properties, 2nd ed. John Wiley and<br />

Sons, New York, NY. 456 p.<br />

Hanski, I., and O. Ovaskainen. 2002. Extinction debt<br />

at extinction thresholds. Conservation Biology 16:666-<br />

673.<br />

Ideker, J. 1999. Wherry mimosa and black mangrove:<br />

<strong>Native</strong>s. <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> of Texas News 17<br />

(5):8 [Online]. Available: http://www.npsot.org/<br />

NPSOTNews/ [June 8, 2009].<br />

Jahrsdoerfer, S. E. & D. M. Leslie. 1988. Tamaulipan<br />

brushland of the Rio Grande Valley of south Texas:<br />

Description, human impacts, and management options.<br />

U.S. Fish and Wildlife Service, Biol. Rep. 88(36), xiii +<br />

63 p.<br />

McLendon, T. 1991. Preliminary description of the<br />

vegetation of south Texas exclusive of coastal saline<br />

zones. Texas Journal of Science 43:13-32.<br />

Moore, M. J. and R. K. Jansen. 2007. Origins and<br />

biogeography of gypsophily in the Chihuahuan Desert<br />

plant group Tiquilia subg. Eddya (Boraginaceae). Systematic<br />

Botany 32(2): 392–414.<br />

NatureServe. 2002. NatureServe Element Occurrence<br />

data standard. Version published February 6, 2002<br />

[Online]. Available: http://www.natureserve.org/<br />

prodServices/eodata.jsp [June 5, 2009].<br />

NatureServe. 2009. NatureServe Explorer: An online<br />

encyclopedia of life. Version 7.0. NatureServe, Arlington,<br />

Virginia. [Online]. Available: http://www.nature<br />

serve.org/explorer [June 5, 2009].<br />

NOAA (National Oceanic and Atmospheric Administration).<br />

2009. National Climatic Data Center, Asheville,<br />

NC [Online]. Available: http://www.ncdc.noaa.<br />

gov/oa/ncdc.html [June 5, 2009].<br />

NRCS (Natural Resources Conservation Service)<br />

Soil Survey Staff. 2009. United States Department of<br />

Agriculture. Web Soil Survey. [Online]. Available:<br />

http://websoilsurvey.nrcs.usda.gov/ [April 30, 2009].<br />

Poole, J. M. 1989. Status report on Lesquerella thamnophila.<br />

Report prepared for U.S. Fish & Wildlife Service,<br />

Albuquerque. 21p. + figs.<br />

Poole, J. M., W. R. Carr, D. M. Price, and J. R.<br />

Singhurst. 2007. Rare <strong>Plant</strong>s of Texas. Texas A&M<br />

University Press, College Station, Texas. 640p.<br />

Preston, R.D. 2009. The Yegua-Jackson aquifer.<br />

Chapter 3, p. 51-59 in: Mace, R.E. et al., eds. Aquifers<br />

of the Gulf Coast of Texas. Texas Water Development<br />

Board Report 365, Austin, Texas. [Online]. Available:<br />

http://www.twdb.state.tx.us/publications/reports/<br />

GroundWaterReports/GWReports/GWreports.asp [June<br />

9, 2009].<br />

Rollins, R.C. and E.A. Shaw. 1973. The genus<br />

Lesquerella (Cruciferae) in North America. Harvard<br />

University Press, Cambridge, Mass.<br />

Sands, J. P., L. A Brennan, F. Hernández, W. P.<br />

Kuvlesky, J. F. Gallagher, D. C. Ruthven, and J.E.<br />

Pittman. 2009. Impacts of buffelgrass (Pennisetum<br />

ciliare) on a forb community in South Texas. Invasive<br />

<strong>Plant</strong> Science and Management, 2(2):130-140.<br />

Sternberg, M. A. 2005. Zapata bladderpod<br />

(Lesquerella thamnophila Roll. & Shaw): its status and<br />

association with other plants. Texas Journal of Science<br />

57(1):75-86.<br />

Thompson, C.M., R.R. Sanders, and D. Williams.<br />

1972. Soil survey of Starr County, Texas. Soil Conservation<br />

Service, United States Department of Agriculture,<br />

in cooperation with the Texas Agricultural Experiment<br />

Station. 88pp. + figs.<br />

USDA NRCS. <strong>2012</strong>. The PLANTS Database (http://<br />

plants.usda.gov, 21 February <strong>2012</strong>). National <strong>Plant</strong><br />

Data Team, Greensboro, NC.<br />

USFWS (U.S. Fish and Wildlife Service). 2004. Zapata<br />

Bladderpod (Lesquerella thamnophila) Recovery<br />

Plan. U.S. Fish and Wildlife Service, Albuquerque, New<br />

Mexico. i-vii + 30 pp., Appendices A-B. [Online].<br />

Available: http://ecos.fws.gov/speciesProfile/profile/<br />

speciesProfile.action?spcode=Q13L [May 27, 2009].<br />

Wu, X.B. and F.E. Smeins. 1999. Multiple-Scale<br />

Habitat Models of Rare <strong>Plant</strong>s: Model Development and<br />

Evaluation. Report Submitted to Texas Department of<br />

Transportation, Pharr and Austin, TX.<br />

188


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Intraspecific Cytotype Variation and Conservation:<br />

An Example from Phlox (Polemoniaceae)<br />

Shannon D. Fehlberg<br />

Desert Botanical Garden, Phoenix, AZ<br />

and Carolyn J. Ferguson,<br />

Herbarium and Division of Biology, Kansas State University, Manhattan, KS<br />

Abstract. Information on genetic structure in rare plants, such as patterns of genetic diversity, differentiation and<br />

gene flow, is useful when planning management strategies for conservation. However, few studies of genetic structure<br />

in rare plants include an investigation of intraspecific cytotype variation. A number of reviews have suggested<br />

that cytotype variation, or variation in chromosome structure or number, may be more widespread in natural populations<br />

than previously thought. A recent review of federally listed plant species found that 75% of species belonged to<br />

genera exhibiting both interspecific and intraspecific cytotype variation. Cytotype variation among populations of<br />

rare plants raises intriguing questions about the origin(s), extent, evolutionary relationships, and ecological differentiation<br />

of various cytotypes. While traditional cytological methods may not be feasible for population level sampling,<br />

advances in the accuracy and cost of flow cytometry now make examination of intraspecific cytotype variation possible.<br />

We initiated an investigation of cytotype variation in the genus Phlox (Polemoniaceae) using flow cytometry.<br />

Phlox comprises ca. 65 species in North America and includes many poorly studied endemic taxa in the western<br />

United States. Our results indicate noteworthy variation in chromosome numbers (ploidy level) across a subset of<br />

western taxa. A detailed study of two endemic species of conservation concern, P. amabilis and P. woodhousei, revealed<br />

that these species are made up of diploid, tetraploid and hexaploid populations. Ongoing analyses suggest that<br />

these populations are spatially, ecologically, and genetically differentiated. These results caution against the common<br />

assumption of homoploidy of species based on limited data and indicate the value of incorporating an understanding<br />

of cytotype variation into conservation biology studies.<br />

The interdisciplinary field of conservation biology<br />

aims to provide the knowledge and tools necessary for<br />

the long-term preservation of biodiversity. Studies of<br />

population genetics are one important source of information<br />

for conservation biology. These studies can provide<br />

basic information about populations of rare plants<br />

such as levels of genetic variation, the distribution of<br />

genetic variation within and among populations, the degree<br />

of population fragmentation and isolation, the patterns<br />

of historical and contemporary gene flow, the levels<br />

of inbreeding, the effective population size, the presence<br />

of taxonomic distinctiveness, and the occurrence of<br />

hybridization (Booy et al. 2000; Ellis and Burke 2007;<br />

Murray and Young 2001). Ultimately, this basic information<br />

can be combined with other sources of data to<br />

form more effective conservation strategies.<br />

One component of genetic variation that is often<br />

overlooked in studies of rare plants is intraspecific cytotype<br />

variation (De Lange et al. 2008; Severns and Liston<br />

2008). Intraspecific cytotype variation ranges from<br />

chromosomal inversions and translocations to variation<br />

in the number of whole genomes present (polyploidy).<br />

Recent reviews have suggested that intraspecific cytotype<br />

variation may be more widespread in natural plant<br />

populations than previously thought (Soltis et al. 2007;<br />

Suda et al. 2007). It may also be widespread in populations<br />

of rare, threatened and endangered plants. A survey<br />

of cytotype variation in 416 US federally listed<br />

plant taxa found that cytotype data were available for<br />

only 182 of these taxa (44%). Of these 182 taxa, 158<br />

belonged to genera with interspecific cytotype variation<br />

(87%; New World congeners with aneuploidy or polyploidy),<br />

and 121 belonged to genera with at least one<br />

taxon showing intraspecific cytotype variation (66%;<br />

Severns and Liston 2008). Intraspecific cytotype variation<br />

can be important when planning conservation<br />

strategies because it has the potential to affect patterns<br />

of genetic diversity and gene flow. Knowledge of such<br />

variation may also be important for clarifying taxonomy,<br />

identifying unique evolutionary lineages (perhaps<br />

accompanied by ecological differentiation), and determining<br />

appropriate populations for reintroduction or<br />

augmentation (De Lange et al. 2008; Murray and Young<br />

2001; Severns and Liston 2008).<br />

Recent advances in the application of flow cytometry<br />

to evolutionary and population biology studies now<br />

make the assessment of cytotype variation across large<br />

taxonomic and spatial scales possible (Kron et al. 2007).<br />

189


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Flow cytometry is a method that quantifies nuclear<br />

DNA content by measuring the relative fluorescence of<br />

isolated nuclei that have been stained with a fluorescent<br />

dye. This method offers several advantages over traditional<br />

cytological methods including: 1) sample preparation<br />

and processing is relatively easy and rapid (allowing<br />

for large sample sizes), 2) sample material does not<br />

need to be actively dividing (a variety of tissues can be<br />

used, including dried and frozen material), and 3) sampling<br />

is relatively non-destructive (enabling studies of<br />

sensitive plants; Kron et al. 2007; Suda et al. 2007).<br />

Thus, flow cytometry provides a practical means by<br />

which intraspecific population level cytotype data can<br />

be gathered.<br />

The genus Phlox has been an important system for<br />

plant evolutionary studies of polyploidy, hybridization,<br />

and ecology (Ferguson and Jansen 2002; Ferguson et al.<br />

1999; Grant 1959; Levin and Schaal 1970; Levin and<br />

Smith 1966; Wherry 1955). Phlox comprises ca. 65 species<br />

of annual and perennial herbs distributed predominantly<br />

in North America with a center of diversity in the<br />

western United States. We are using flow cytometry to<br />

study cytotype variation across this genus. The present<br />

study focuses on a broad sample of western, upright perennial<br />

Phlox species with special emphasis on two endemic<br />

species of conservation concern found in coniferous<br />

forests in Arizona and New Mexico, P. amabilis and<br />

P. woodhousei. These two species are closely related<br />

(Ferguson et al. <strong>2012</strong>) and very similar based on gross<br />

morphology, sharing a characteristic woody-based upright<br />

perennial growth form, thick linear leaves and<br />

notched petals. However, their geographic distributions<br />

do not overlap, and they are readily distinguished by<br />

differences in style length and stigma placement relative<br />

to anther position. These taxa have been variously classified<br />

as distinct species or conspecifics (Cronquist et al.<br />

1984; Wherry 1955). Our objectives for the present<br />

study were to examine 1) taxonomic and large-scale<br />

spatial patterns of cytotype variation across a subset of<br />

western Phlox taxa and 2) fine-scale spatial patterns of<br />

cytotype variation within and among populations of P.<br />

amabilis and P. woodhousei.<br />

METHODS<br />

The present study includes samples from seven species<br />

of upright, perennial Phlox from Arizona, California,<br />

Colorado, Idaho, Montana, Nevada, <strong>Utah</strong> and Wyoming<br />

(Table 1). For determination of cytotype, several<br />

leaves were collected from one to five individual plants<br />

at each sampling location for each taxon and stored on<br />

ice until nuclear extraction. Voucher specimens for each<br />

population were deposited at the Kansas State University<br />

Herbarium (KSC).<br />

We determined the cytotype of each sample of Phlox<br />

using flow cytometry. Flow cytometry measures nuclear<br />

DNA content, which can then be interpreted in terms of<br />

ploidy level, especially when closely related taxa are<br />

studied and when knowledge of ploidy level is independently<br />

assessed through conventional chromosome<br />

counts (Suda et al. 2007; Halverson et al. 2008). For<br />

each sample of Phlox, we placed 100-300 mg of chilled<br />

leaf tissue into a petri dish with 1.5 ml of chopping<br />

buffer, modified from Bino and others (1993) as described<br />

by Davison and others (2007). We chopped the<br />

leaves finely with a new razor blade and filtered the resulting<br />

liquid through a 30 µm filter into a microcentrifuge<br />

tube. Tubes were centrifuged at 500 x g for 7 min,<br />

the supernatant removed, the pellet re-suspended in 700<br />

µl propidium iodide staining solution (50 mg/ml;<br />

BioSure), and 2µl of chicken erythrocyte nuclei singlets<br />

added (CEN internal standard; BioSure). Samples were<br />

protected from light and stored on ice for at least 30 min<br />

before analysis on a Becton Dickinson FACS Calibur<br />

flow cytometer at the Kansas State University Flow Cytometry<br />

Facility. The amount of fluorescence was measured<br />

for ~10,000 nuclei per sample. Resulting histograms<br />

were visually inspected for the presence of clear<br />

nuclear populations from the Phlox sample and the CEN<br />

internal standard, and mean peak values were calculated<br />

using the program Cell Quest (Becton Dickinson). Nuclear<br />

DNA content was calculated as the Phlox sample<br />

mean peak value divided by the CEN internal standard<br />

mean peak value multiplied by the 2C-value of the CEN<br />

internal standard (2.5 pg; Dolezel and Bartos 2005).<br />

Ploidy level was inferred for each sample based on the<br />

calculated DNA content. Inferred ploidy level was<br />

linked with chromosome count data for several samples.<br />

RESULTS AND DISCUSSION<br />

A total of 140 samples from seven species of Phlox<br />

collected from 63 locations were assessed for cytotype<br />

variation (Table 1). When cytotype was measured in<br />

multiple individuals from a single location, average nuclear<br />

DNA content was calculated. We did not detect<br />

any cytotype variation within populations based on our<br />

limited sampling. The results from flow cytometry were<br />

interpreted as measures of ploidy level rather than absolute<br />

measures of DNA content (see Suda et al. 2007).<br />

Results from chromosome counts confirmed ploidy levels<br />

of 2x, 4x, and 6x (for five samples; Table 1).<br />

Cytotype varied throughout the species studied and<br />

appeared to reflect both taxonomy and geography<br />

(Table 1). Populations of P. caryophylla and P. cluteana<br />

were diploid, while populations of P. aculeata were<br />

tetraploid; all of these taxa are fairly narrow endemics.<br />

In general, populations of the wide-ranging P. longifolia<br />

and P. stansburyi were geographically structured, with<br />

190


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. Samples included in this study. Voucher information, collection locality, number of individuals<br />

cytotyped (N), DNA content and inferred cytotype are noted. Taxon recognition for P. stansburyi follows<br />

Ferguson and others (<strong>2012</strong>); intraspecific classification for P. longifolia (including P. longifolia and P. viridis<br />

sensu Wherry [1955]) is under ongoing investigation and is not proposed here.<br />

Taxon Voucher County State Location Detail<br />

N<br />

DNA content<br />

(pg)<br />

P. aculeata Sidells 116 Twin Falls Co. ID 1 14.65 4X<br />

P. aculeata CF 789 Butte Co. ID 3 15.43 4X<br />

P. aculeata CF 787 Bonneville Co. ID 2 18.67 4X<br />

P. caryophylla CF 653 Archuleta Co. CO 1 7.61 2X<br />

P. cluteana CF 651 Apache Co. AZ 1 8.14 2X<br />

P. cluteana SDF 51008-2 Apache Co. AZ 3 8.84 2X<br />

P. longifolia SDF 42908-3 Navajo Co. AZ 3 8.37 2X<br />

P. longifolia SDF 43008-1 Coconino Co. AZ 3 8.60 2X<br />

P. longifolia SDF 42908-1 Navajo Co. AZ 3 8.74 2X<br />

P. longifolia SDF 51008-1 Apache Co. AZ 3 8.88 2X<br />

P. longifolia SDF 50408-1 Coconino Co. AZ 3 8.89 2X<br />

P. longifolia SDF 50908-1 Coconino Co. AZ 3 9.17 2X<br />

P. longifolia SDF 50708-1 Mojave Co. AZ 3 9.27 2X<br />

P. longifolia SDF 50808-1 Mojave Co. AZ 3 16.79 4X<br />

P. longifolia SDF 50808-2 Mojave Co. AZ 4 17.28 4X<br />

P. longifolia CF 728 Eagle Co. CO 1 15.25 4X<br />

P. longifolia CF 727 Mesa Co. CO 1 16.71 4X<br />

P. longifolia CF 788 Butte Co. ID 3 7.60 2X<br />

P. longifolia CF 806 Lemhi Co. ID 1 14.27 4X<br />

P. longifolia CF 603 Clark Co. ID 1 14.67 4X<br />

P. longifolia CF 792 Butte Co. ID 3 15.25 4X<br />

P. longifolia CF 589 Custer Co. ID 1 15.41 4X<br />

P. longifolia CF 794 Custer Co. ID 3 15.67 4X<br />

P. longifolia CF 786 Jefferson Co. ID 2 18.67 4X<br />

P. longifolia CF 601 Beaverhead Co. MT 1 13.63 4X<br />

P. longifolia CF 710 White Pine Co. NV 1 17.11 4X<br />

P. longifolia CF 724 Beaver Co. UT 1 8.38 2X<br />

P. longifolia CF 702 Sevier Co. UT 1 15.00 4X<br />

P. longifolia CF 611 Summit Co. UT 1 21.90 4X<br />

P. longifolia CF 734 Carbon Co. WY 1 13.98 4X a<br />

P. longifolia CF 215 Sublette Co. WY 1 14.65 4X<br />

Cytotype<br />

191


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. continued<br />

Taxon Voucher County State Location Detail<br />

192<br />

N<br />

DNA content<br />

(pg)<br />

P. longifolia CF 617 Beaver Co. UT 1 15.25 4X<br />

P. longifolia CF 606 Lincoln Co. WY 1 14.51 4X<br />

P. stansburyi var.<br />

brevifolia<br />

P. stansburyi var.<br />

brevifolia<br />

P. stansburyi var.<br />

brevifolia<br />

P. stansburyi var.<br />

brevifolia<br />

P. stansburyi var.<br />

brevifolia<br />

P. stansburyi ssp.<br />

stansburyi<br />

P. stansburyi ssp.<br />

stansburyi<br />

SDF 50208-2 Yavapi Co. AZ 4 8.60 2X<br />

SDF 50308-2 Yavapi Co. AZ 3 9.18 2X<br />

SDF 50508-3 Coconino Co. AZ 4 9.63 2X<br />

CF 630 Mono Co. CA 1 8.31 2X<br />

CF 718 Kane Co. UT 1 10.49 2X<br />

SDF 42808-1 Cochise Co. AZ 3 8.18 2X<br />

SCS 18 Lander Co. NV 1 15.85 4X<br />

P. stansburyi ssp. superba SDF 50908-2 Coconino Co. AZ 3 18.86 4X<br />

P. stansburyi ssp. superba CF 634 Inyo Co. CA 1 10.54 2X<br />

P. stansburyi ssp. superba CF 704 Nye Co. NV 1 7.31 2X<br />

P. stansburyi ssp. superba CF 705 Nye Co. NV 1 12.01 2X<br />

P. stansburyi ssp. superba CF 713 Nye Co. NV 1 15.69 4X<br />

P. stansburyi ssp. superba CF 707 White Pine Co. NV 1 18.87 4X<br />

P. amabilis CF 780 Yavapai Co. AZ Camp Woods 1 8.28 2X<br />

P. amabilis CF 775 / SDF<br />

51507-2<br />

Yavapai Co. AZ Thumb Butte 3 8.41 2X b<br />

P. amabilis SDF 51707-1 Mojave Co. AZ Black Rock 1 8.81 2X<br />

P. amabilis SDF 51807-2 /<br />

50708-4<br />

Mojave Co. AZ Death Valley<br />

Spring<br />

5 16.92 4X<br />

P. amabilis SDF 50208-4 Yavapai Co. AZ Watson Lake 2 17.58 4X<br />

P. amabilis SDF 50308-1 Yavapai Co. AZ Mingus Mtn 4 24.52 6X b<br />

P. amabilis SDF 51407-2 Coconino Co. AZ Hobble Mtn 3 24.54 6X<br />

P. amabilis SDF 50508-1 Coconino Co. AZ Kaibab Lake 3 26.36 6X<br />

P. woodhousei CF 770 Coconino Co. AZ Oakcreek 1 8.81 2X<br />

P. woodhousei SDF 50108-3 Coconino Co. AZ Stoneman 3 8.89 2X b<br />

P. woodhousei SDF 51407-1 Coconino Co. AZ Bill Williams 1 9.02 2X<br />

P. woodhousei SDF 50807-1 Catron Co. NM Reserve 1 16.50 4X<br />

P. woodhousei SDF 50108-1 Gila Co. AZ McFadden<br />

Peak<br />

5 16.89 4X<br />

Cytotype


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 1. continued<br />

Taxon Voucher County State Location Detail<br />

N<br />

DNA content<br />

(pg)<br />

P. woodhousei SDF 42908-2 Navajo Co. AZ Showlow 3 17.76 4X<br />

P. woodhousei SDF 50108-2 Coconino Co. AZ Strawberry 6 18.13 4X b<br />

P. woodhousei SDF 43008-2 Gila Co. AZ Sharp Creek 8 20.06 4X<br />

P. woodhousei SDF 51107-2 Gila Co. AZ Sierra Ancha 1 27.16 6X<br />

a inferred ploidy level supported by published mitotic chromosome count from the same population (Löve, 1971; counted by<br />

Daniel J. Crawford; a voucher specimen from New York Botanical Garden [Crawford 76, June 1970] was also consulted).<br />

b inferred ploidy level supported by meitoic chromosome counts conducted in laboratories of the authors.<br />

Cytotype<br />

tetraploid populations located in the northern portion of<br />

the range and diploid populations located in the southern<br />

portion of the range. Cytotype in some populations<br />

of P. stansburyi subsp. superba could not be reliably<br />

determined due to wide variation in DNA content, and<br />

further study of this taxon is needed. This preliminary<br />

survey indicates that cytotype variation may be a useful<br />

character for clarifying historically complicated tax-<br />

onomic divisions (as suggested by Kron et al. 2007)<br />

among these Phlox species.<br />

An in-depth survey of cytotype variation in P.<br />

amabilis and P. woodhousei revealed that these species<br />

were made up of diploid, tetraploid and hexaploid populations<br />

and that some cytotypes were restricted to specific<br />

portions of the range (Table 1; Figure 1). For example,<br />

most populations of P. woodhousei were associ-<br />

Figure 1. Map showing the locations and ploidy levels of 18 populations of Phlox amabilis and P. woodhousei sampled<br />

for this study.<br />

193


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

ated with the Mogollon Rim formation, but diploid<br />

populations appeared to be restricted to the western end<br />

of the rim. Hexaploid populations of P. amabilis were<br />

restricted to the easternmost portion of the range and<br />

appeared to be associated with magnesium rich igneous<br />

rock formations (data not shown). It is possible that<br />

such differences in distribution among cytotypes within<br />

species reflect ecological differentiation, which could<br />

result in local adaptation (Schonswetter et al. 2007;<br />

Buggs and Pannell 2007; Paun et al. 2007). Preliminary<br />

analysis of genetic variation in P. amabilis and P. woodhousei<br />

also supports differentiation among cytotypes<br />

(data not shown). We are continuing to investigate this<br />

unexpected variation in cytotype in both species using<br />

insights from genetics and ecology.<br />

CONCLUSIONS<br />

The results of this survey of cytotype variation<br />

among these western Phlox taxa demonstrate the value<br />

of such knowledge for the study of plant diversity, evolution,<br />

and conservation. The use of flow cytometry allowed<br />

us to gather data easily and rapidly on cytotype<br />

variation across expanded sample sizes that included<br />

multiple species from multiple locations throughout<br />

their ranges as well as population level sampling in species<br />

of conservation interest. We found that cytotype<br />

was variable not only among species but also among<br />

populations within species. This variation appears to be<br />

related to taxonomy, geography, and possibly ecology.<br />

Continued studies of cytotype variation in Phlox will<br />

provide valuable data for resolving taxonomic divisions<br />

as well as insight into the evolutionary diversification of<br />

this group. The detailed analysis of cytotype variation in<br />

P. amabilis and P. woodhousei, combined with results<br />

from ongoing population genetic work, suggest the presence<br />

of unique evolutionary lineages and ecological differentiation,<br />

both of which are important when planning<br />

conservation strategies. Taken together with recent reviews,<br />

these results emphasize the value of incorporating<br />

an understanding of cytotype variation into conservation<br />

biology studies.<br />

ACKNOWLEDGEMENTS<br />

We gratefully acknowledge support from the KSU<br />

Ecological Genomics Institute, the KSU Center for the<br />

Understanding of Origins and NSF DEB-0089656. We<br />

thank Thelma Green and Maureen Ty for technical assistance,<br />

Jerry Davison and Suzi Strakosh for assistance<br />

with the initial stages of flow cytometry, and Mark<br />

Mayfield and Ed Turcotte for unpublished chromosome<br />

count data. We thank the New York Botanical Garden<br />

(NY) for a loan of herbarium specimens. We thank<br />

Montezuma Castle National Monument (MOCA-2008-<br />

SCI-0004), Grand Canyon National Park (GRCA-2008-<br />

SCI-0015), the National Forest Service (various permits,<br />

Regions 1-6), the California BLM, and the Navajo Nation<br />

(Permits 414 and 051402-062) for permission to<br />

collect plants. This is publication 09-372-A of the Kansas<br />

Agricultural Experiment Station.<br />

LITERATURE CITED<br />

Bino, R.J., S. Lanteri, H.A. Verhoeven, and H.L.<br />

Kraak. 1993. Flow cytometric determination of nuclear<br />

replication stages in seed tissues. Annals of Botany 72:<br />

181-187.<br />

Booy, G., R.J.J. Hendriks, M.J.M. Smulders, J.M.<br />

Van Groenendael, and B. Vosman. 2000. Genetic diversity<br />

and the survival of populations. <strong>Plant</strong> Biology 2:<br />

379-395.<br />

Buggs, R.J.A. and J.R. Pannell. 2007. Ecological<br />

differentiation and diploid superiority across a moving<br />

ploidy contact zone. Evolution 61: 125-140.<br />

Cronquist A., A.H. Holmgren, N.H. Holmgren, J.L.<br />

Reveal, and P.K. Holmgren. 1984. Intermountain Flora,<br />

Vol. 4. New York Botanical Garden, Bronx, New York.<br />

Davison, J., A. Tyagi, and L. Comai. 2007. Largescale<br />

polymorphism of heterochromatic repeats in the<br />

DNA of Arabidopsis thaliana. BMC <strong>Plant</strong> Biology 7:<br />

44.<br />

De Lange, P.J., P.B. Heenan, D.J. Keeling, B.G.<br />

Murray, R. Smissen, and W.R. Sykes. 2008. Biosystematics<br />

and conservation: a case study with two enigmatic<br />

and uncommon species of Crassula from New<br />

Zealand. Annals of Botany 101: 881-99.<br />

Dolezel, J. and J. Bartos. 2005. <strong>Plant</strong> DNA flow cytometry<br />

and estimation of nuclear genome size. Annals<br />

of Botany 95: 99-110.<br />

Ellis, J.R. and J.M. Burke. 2007. EST-SSRs as a resource<br />

for population genetic analyses. Heredity 99:<br />

125-132.<br />

Ferguson, C.J. and R.K. Jansen. 2002. A chloroplast<br />

DNA phylogeny of eastern Phlox (Polemoniaceae): implications<br />

of congruence and incongruence with the ITS<br />

phylogeny. American Journal of Botany 89: 1324-1335.<br />

Ferguson, C.J., F. Krämer, and R.K. Jansen. 1999.<br />

Relationships of eastern North American Phlox<br />

(Polemoniaceae) based on ITS sequence data. Systematic<br />

Botany 24: 616-631.<br />

Ferguson, C.J., S.C. Strakosh, and R. Patterson.<br />

<strong>2012</strong>. Phlox. Pp. 1068-1071 In: B. G. Baldwin, D.H.<br />

Goldman, D.J. Keil, R. Patterson, T.J. Rosatti, and D.H.<br />

Wilken, eds. The Jepson Manual, Vascular <strong>Plant</strong>s of<br />

California, second edition. University of California<br />

Press, Berkeley.<br />

Grant, V. 1959. Natural history of the Phlox family:<br />

systematic botany. Martinus Nijhoff, The Hague, Netherlands.<br />

Halverson, K., S.B. Heard, J.D. Nason, and J.O.<br />

Stireman III. 2008. Origins, distribution, and local cooccurrence<br />

of polyploidy cytotypes in Solidago altis-<br />

194


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

sima (Asteraceae). American Journal of Botany 95: 50-<br />

58.<br />

Kron, P., J. Suda, and B.C. Husband. 2007. Applications<br />

of flow cytometry to evolutionary and population<br />

biology. Annual Review of Ecology, Evolution, and<br />

Systematics 38: 847-876.<br />

Levin, D.A. and B.A. Schaal. 1970. Reticulate evolution<br />

in Phlox as seen through protein electrophoresis.<br />

American Journal of Botany 57: 977-987.<br />

Levin, D.A. and D.M. Smith. 1966. Hybridization<br />

and evolution in the Phlox pilosa complex. The American<br />

Naturalist 100: 289-301.<br />

Löve, A. 1971. IOPB chromosome number reports<br />

XXXI. Taxon 20:157-160.<br />

Murray, B.G. and A.G. Young. 2001. Widespread<br />

chromosome variation in the endangered grassland forb<br />

Rutidosis leptorrhynchoides F. Muell. (Asteraceae: Gnaphalieae).<br />

Annals of Botany 87: 83-90.<br />

Paun, O., F.M. Fay, D.E. Soltis, and M.W. Chase.<br />

2007. Genetic and epigenetic alterations after hybridization<br />

and genome doubling. Taxon 56: 649-656.<br />

Schonswetter, P., M. Lachmayer, C. Lettner, D. Prehsler,<br />

S. Rechnitzer, D.S. Reich, M. Sonnleitner, I.<br />

Wagner, K. Huelber, G.M. Schneeweiss, P. Travnicek,<br />

and J. Suda. 2007. Sympatric diploid and hexaploid cytotypes<br />

of Senecio carniolicus (Asteraceae) in the Eastern<br />

Alps are separated along an altitudinal gradient.<br />

Journal of <strong>Plant</strong> Research 120: 721-725.<br />

Severns, P.M. and A. Liston. 2008. Intraspecific<br />

chromosome number variation: a neglected threat to the<br />

conservation of rare plants. Conservation Biology 22:<br />

1641-1647.<br />

Soltis, D.E., P.S. Soltis, D.W. Schemske, J.F. Hancock,<br />

J.N. Thompson, B.C. Husband, and W.S. Judd.<br />

2007. Autopolyploidy in angiosperms: have we grossly<br />

underestimated the number of species? Taxon 56: 13-<br />

30.<br />

Suda, J., P. Kron, B.C. Husband, P. Travnicek. 2007.<br />

Flow cytometry and ploidy: applications in plant systematics,<br />

ecology and evolutionary biology. In: J.<br />

Dolezel, J. Greilhuber, and J. Suda, eds. Flow Cytometry<br />

with <strong>Plant</strong> Cells: Analysis of Genes, Chromosomes<br />

and Genomes. Wiley-VCH, Weinheim. Pp. 103-130.<br />

Wherry, E.T. 1955. The genus Phlox. Morris Arboretum<br />

Monographs 3. Philadelphia, Pennsylvania.<br />

195


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Prioritizing <strong>Plant</strong> Species for Conservation in <strong>Utah</strong>:<br />

Developing the UNPS Rare <strong>Plant</strong> List<br />

Walter Fertig, Chair, <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> Rare <strong>Plant</strong> Committee<br />

Abstract. Rare plant lists are an important tool for identifying and prioritizing species for conservation attention.<br />

Over a dozen systems have been derived for ranking the rarity and conservation priority of plant and animal species,<br />

each differing in emphasis, methods, and biological and anthropogenic criteria. In 2007 I developed a new ranking<br />

protocol for the flora of Wyoming that combines aspects of the NatureServe, International Union for Conservation of<br />

Nature (IUCN), and US Fish and Wildlife Service systems and the classic paper “Seven forms of Rarity” by Deborah<br />

Rabinowitz. The so-called “Wyoming protocol” was adopted by the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>’s Rare <strong>Plant</strong> Committee<br />

to develop an updated rare plant list for <strong>Utah</strong>. In this protocol, species or varieties are assessed using seven qualitative<br />

criteria: <strong>Utah</strong>’s contribution to global distribution, number of populations in the state, number of individuals,<br />

habitat specificity, intrinsic rarity, magnitude of threats, and population trend. Individual criteria are rated on a binary<br />

scale (0 for unthreatened, 1 for at risk) based on expert opinion. Species for which no data are available are scored<br />

“unknown”. The values for each criterion are summed to derive a minimum and maximum potential score for each<br />

taxon. The minimum score is calculated by summing each individual score and treating any unknown criteria as 0.<br />

The maximum score is derived in the same way, except that unknown criteria are given a value of 1. The two summary<br />

scores are averaged to determine a conservation priority rank. Those taxa that are at risk for a large number of<br />

criteria have higher conservation priority ranks than those species that are at risk for only a few criteria. This simple<br />

method allows practitioners to rapidly identify the relatively small subset of species of high or extremely high conservation<br />

priority (those with limited ranges, few populations, low numbers, high habitat specificity, high intrinsic rarity,<br />

high threats, and downward trends) and those species with significant data gaps in need of additional study. Being<br />

able to differentiate among species based on their priority score enables conservationists and managers to better allocate<br />

limited resources to those taxa most in need. The UNPS <strong>Utah</strong> rare plant list developed by the Rare <strong>Plant</strong> Committee<br />

and attendees of a breakout ranking session at the Fifth Southwestern Rare <strong>Plant</strong> Conference is presented in<br />

Appendices 1-4, with modifications adopted at subsequent <strong>Utah</strong> Rare <strong>Plant</strong> meetings from 2010-<strong>2012</strong>.<br />

Experts predict that one-fifth to one-third of all vascular<br />

plant species in the United States are threatened<br />

with local or range-wide extinction (Center for <strong>Plant</strong><br />

Conservation 2000). This number is only likely to increase<br />

as plant habitat becomes increasingly fragmented<br />

and disturbed by development, climate change, or invasion<br />

by non-native weeds. Not all plant species, however,<br />

are equally imperiled. Some species are naturally<br />

rare due to their limited range, high habitat specificity,<br />

or low population size (Rabinowitz 1981), but may not<br />

be in imminent danger because their population trends<br />

are stable or threats are presently low. Because so many<br />

species are potentially vulnerable and conservation resources<br />

(time, funding, and personnel) are nearly always<br />

inadequate, conservation biologists have a dilemma determining<br />

which species should be the highest priority<br />

for attention (Noss and Cooperrider 1994; Regan 2005).<br />

Rare species lists can be an important tool for identifying<br />

and prioritizing those taxa (species, subspecies,<br />

and varieties) most vulnerable to extinction. Over the<br />

past 40 years conservation biologists have proposed<br />

more than a dozen ranking systems for creating state or<br />

national rare species lists (Andelman et al. 2004).<br />

Ranking schemes often differ widely in their emphasis<br />

on inherent rarity, degree of threat, vulnerability of extinction<br />

as well as their scoring methods and overall<br />

complexity and transparency (Akcakaya et al. 2000;<br />

Faber-Langendoen et al. 2009; IUCN 2001; O’Grady et<br />

al. 2004; Rabinowitz 1981; Regan et al. 2004; Spence<br />

<strong>2012</strong>, US Fish and Wildlife Service 1983). These systems<br />

also utilize different criteria for ranking, including<br />

abundance, number of populations, geographic range,<br />

area of occupancy, population trend, intrinsic rarity,<br />

taxonomic distinctiveness, ecological significance,<br />

population viability, habitat condition or degree of fragmentation,<br />

magnitude and imminence of threats, and<br />

number of protected populations (Andelmann et al.<br />

2004; Beissinger et al. 2000; Breininger et al. 1998;<br />

Holsinger 1992; IUCN 2001; Keith 1998; Mace et al.<br />

2008; Millsap et al. 1990; Panjabi et al. 2005; Rabinowitz<br />

1981; Regan et al. 2004; Spence <strong>2012</strong>; US Fish<br />

and Wildlife Service 1983).<br />

Ideally, a ranking system should have a strong biological<br />

basis, recognize the significance of threats and<br />

trends, be easy to apply and update with available information<br />

(while recognizing the importance of data gaps),<br />

and be transparent (Fertig 2011). Each ranking system<br />

has its merits, but none meet all of these criteria. For<br />

196


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

example, the US Fish and Wildlife Service (1983) system<br />

for listing species under the US Endangered Species<br />

Act is heavily weighted towards threats and taxonomic<br />

distinctiveness at the expense of other aspects of rarity<br />

(Master et al. 2000). The IUCN (2001) protocol focuses<br />

chiefly on population size, trends, and likelihood of extinction<br />

but is dependent on quantitative viability data<br />

that are not always available for vascular plants. One<br />

advantage of the IUCN protocol, however, is its recognition<br />

of “data deficient” species (Akçakaya et al. 2000).<br />

Rabinowitz (1981) introduced a simple, but elegant,<br />

binary ranking system using just three components of<br />

rarity: geographic range, abundance, and habitat specificity.<br />

But several additional biological and anthropogenic<br />

criteria (such as threat) were not incorporated,<br />

which limits the suitability of the Rabinowitz system for<br />

prioritizing among different kinds of rare species.<br />

The most widely used ranking protocol today is the<br />

natural heritage system, first developed by The Nature<br />

Conservancy in the 1970s (Master 1991) and now administered<br />

by NatureServe. In this system, full species<br />

or varieties are assigned a conservation rank on a scale<br />

of 1 (critically imperiled) to 5 (demonstrably secure)<br />

across their entire global range (G rank) and at a subregional<br />

scale (state/province, or S rank). Traditionally, G<br />

and S ranks were based on the number of occurrences<br />

(discrete biological populations), abundance, or risk of<br />

extinction as determined by expert opinion (Master et al.<br />

2000). In the past decade, NatureServe protocols have<br />

become more quantitative and consider additional ranking<br />

criteria, including long and short-term trends, area<br />

of occupancy, condition of occurrences, intrinsic rarity,<br />

and threat (Faber-Langendoen et al. 2009; Regan et al.<br />

2004). Unfortunately, the revised NatureServe ranking<br />

protocol has become more complex and less transparent.<br />

Individual ratings are weighted differently, some criteria<br />

are used only conditionally, and scores are tallied by a<br />

“black box” computer algorithim (Faber-Langendoen et<br />

al. 2009).<br />

As part of my doctoral dissertation on plant conservation<br />

in Wyoming (Fertig 2011), I developed a hybrid<br />

ranking protocol by borrowing components of each of<br />

the preceding systems. As a starting point, I adopted<br />

most of the rarity factors from NatureServe (Regan et al.<br />

2004), added the uncertainty components of IUCN<br />

(2001), and included an emphasis on threats from the<br />

US Fish and Wildlife Service (1983). The ranking system<br />

itself is a modification of the qualitative, binary<br />

scoring employed by Rabinowitz (1981), expanded to<br />

include additional criteria. I added a simple scoring<br />

component to classify plant species into six different<br />

rarity classes reflecting each taxon’s overall conservation<br />

priority (Fertig 2009, 2011).<br />

In 2007, I beta-tested the “Wyoming protocol” at the<br />

annual <strong>Utah</strong> rare plant meeting sponsored by the <strong>Utah</strong><br />

<strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> (UNPS) and Red Butte Garden.<br />

Following the meeting, the UNPS state board voted to<br />

reestablish a rare plant committee and charged it with<br />

applying this ranking system to the entire <strong>Utah</strong> vascular<br />

plant flora in order to create a new, prioritized list of<br />

rare plant species for the state. A draft version of the<br />

list was presented at a special session of the Fifth Southwestern<br />

Rare and Endangered <strong>Plant</strong> Conference, held at<br />

the University of <strong>Utah</strong> in March 2009. Based on feedback<br />

from meeting participants and other experts, the<br />

list was revised and published in November 2009<br />

(Fertig 2009). The list has since been updated twice<br />

(Fertig 2010a, <strong>2012</strong>) based on additional input from the<br />

UNPS Rare <strong>Plant</strong> Committee, attendees of the <strong>Society</strong>’s<br />

annual rare plant meeting, and review of new literature.<br />

The purpose of this paper is to briefly describe the<br />

Wyoming ranking protocol and its application to the<br />

flora of <strong>Utah</strong>. Appendices 1-4 include the current lists<br />

of <strong>Utah</strong> plants on the Extremely High Priority, High<br />

Priority, Watch, and Need Data lists. The paper concludes<br />

with a comparison of the current list to previous<br />

rare plant lists for <strong>Utah</strong> and a discussion of additional<br />

applications and future directions.<br />

METHODS<br />

Ranking Criteria<br />

The Wyoming protocol is based on seven biological<br />

and anthropogenic factors that influence the conservation<br />

priority of a vascular plant species. These criteria<br />

are:<br />

1. Geographic range. Geographic range takes into<br />

account the state’s contribution to the total global distribution<br />

of a species. Six geographic range categories are<br />

recognized (Table 1). Local and regional endemics<br />

have highly restricted global distributions, ranging from<br />

single populations covering a few acres to less than<br />

250,000 km 2 (an area about the size of the state of Wyoming).<br />

Widespread species, defined as occupying a<br />

global range in excess of 250,000 km 2 , can still be considered<br />

rare if state populations are widely isolated from<br />

the core of the species’ range (disjunct) or are at its very<br />

edge (peripheral). A small number of plant species may<br />

occur widely but are limited to small, often scattered or<br />

discontinuous habitats and occupy less than 5% of the<br />

state (sparse). Species that are introduced to the state are<br />

not included in the rankings.<br />

2. <strong>Number</strong> of Populations. This criterion is based on<br />

the number of extant populations of a species within<br />

<strong>Utah</strong> (occurrences outside the state are not considered).<br />

Populations are defined as aggregations of individual<br />

plants within a specific geographic area that are separated<br />

from other populations by a physical barrier, extensive<br />

area of unsuitable habitat, or sufficient distance<br />

to prevent gene flow (usually about 1-2 km). The num-<br />

197


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Table 1. Scores for Ranking Factors.<br />

Ranking Factor Category or Condition Points<br />

1. Geographic Range<br />

(only taxa native to the<br />

state are scored)<br />

2. <strong>Number</strong> of<br />

Populations<br />

3. Abundance<br />

Local endemic (global range less than 16,500 km 2 or about 1 degree of latitude x 2<br />

degrees of longitude)<br />

Regional endemic (global range covering 16,501-250,000 km 2 or an area about the<br />

size of Wyoming)<br />

Disjunct (globally widespread but state population is isolated from the main contiguous<br />

range of the species by a gap of more than 800 km)<br />

Peripheral (globally widespread but state population is at the margin of its continuous<br />

range and occupies less than 5% of the state’s area near state boundary)<br />

Sparse (globally widespread, but distribution patchy and discontinuous in the state<br />

and covering less than 5% of the state’s area)<br />

Widespread (occurs widely across North America [covering more than 250,000<br />

km 2 ] and across the state [occupying well over 5% of the area])<br />

Unknown 0-1<br />

Low (fewer than 25 extant populations in state) 1<br />

Medium to High (25 or more extant populations in state) 0<br />

Unknown 0-1<br />

Low (depends on life history of species, but typically less than 30,000 individuals<br />

for perennials [higher numbers allowable for annuals] or occupying an area of less<br />

than 3000 acres in state)<br />

Medium to High (known from well over 30,000 individuals for perennials or occupying<br />

an area greater than 3000 acres in state)<br />

Unknown 0-1<br />

4. Habitat Specificity High (“Specialist” restricted to one or a few specialized geologic substrates, soil<br />

types, or vegetation types)<br />

Medium to Low (“Generalist” found in numerous geologic substrates, soil types, or<br />

vegetation types)<br />

Unknown 0-1<br />

5. Intrinsic Rarity High (unusual life history, dependence on rare or specialized pollinators, poor dispersal,<br />

low fecundity, poor seedling survival, etc.)<br />

6. Magnitude and<br />

Imminence of Threats<br />

Medium to Low (no unusual life history or biological attributes limiting establishment<br />

or persistence<br />

Unknown 0-1<br />

High (current or foreseeable threats significant or broad in scale or scope 1<br />

Medium to Low (threats minimal or limited to small percentage of populations<br />

now or in the foreseeable future)<br />

Unknown 0-1<br />

7. Population Trend Decreasing (short to long-term decline in number, size, or vigor of populations) 1<br />

Increasing, stable, or oscillating around a mean 0<br />

Unknown 0-1<br />

198<br />

2<br />

1<br />

1<br />

1<br />

1<br />

0<br />

1<br />

0<br />

1<br />

0<br />

1<br />

0<br />

0


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

ber of populations is not necessarily equivalent to the<br />

number of collections of a species, especially if these<br />

collections are from the same general area.<br />

3. Abundance. Abundance refers to the number of<br />

individual plants known statewide. If census data are<br />

lacking, abundance can be approximated from the relative<br />

dominance of a species within its area of occupied<br />

habitat.<br />

4. Habitat Specificity. This factor assesses the degree<br />

to which a species is a habitat specialist restricted<br />

to a particular soil or geologic substrate (edaphic endemics)<br />

or vegetation type, or is a generalist found in a<br />

wide variety of substrates or plant communities.<br />

5. Intrinsic Rarity. Analogous to habitat specificity,<br />

intrinsic rarity addresses those attributes of a species’<br />

life history that may limit its establishment or persistence.<br />

Examples include low fecundity, poor dispersal,<br />

low seedling survival, low genetic diversity, or dependence<br />

on specialized pollinators.<br />

6. Magnitude and Imminence of Threats. This criterion<br />

assesses the scope, severity, and immediacy of current<br />

or future negative impacts on a species. Potential<br />

threats include habitat destruction, over-collection, herbivory,<br />

trampling or soil compaction from recreation, or<br />

competition from invasive plants.<br />

7. Population Trend. Trend is the change in population<br />

size, extent, and vigor over time.<br />

Assigning scores to each criterion<br />

Following the model of Rabinowitz (1981), six of the<br />

seven preceding criteria are scored using a binary rating<br />

(high/low or increasing/decreasing). A score of 1 is assigned<br />

to those conditions that make a species highly<br />

vulnerable to extinction or extirpation, while a score of<br />

0 is given for conditions that only moderately impact or<br />

do not adversely affect a species’ persistence in the<br />

state. The only exception is geographic range in which<br />

three scores are possible (0, 1, or 2) to allow greater<br />

weighting of locally endemic species. If there is insufficient<br />

data to rate a specific criterion, or available information<br />

is inconclusive, a value of “U” (unknown) may<br />

be assigned. Scores for individual criteria are shown in<br />

Table 1.<br />

Scoring is based on a review of pertinent literature,<br />

specimen databases, and expert knowledge and should<br />

be supported by corroborating data. Scores can be tabulated<br />

in a spreadsheet or in a simple data form (see Table<br />

2 for an example).<br />

Determining conservation priority<br />

Once the ranking form or table is completed, the individual<br />

scores for each of the seven ranking factors are<br />

summed to derive both a minimum and potential (maximum)<br />

score (Table 2). These scores can range from 0<br />

to 8. The minimum score includes only those factors for<br />

which information is available, with any unknowns<br />

scored as 0. The potential score includes the same values<br />

but with unknowns given a “worst case” score of 1.<br />

The minimum and potential scores are then averaged<br />

(with the sum rounded down) to derive an overall score<br />

(Table 2).<br />

The final score can be used to assign each species to<br />

one of the following six conservation priority categories:<br />

Extremely High (7 or 8 points): species at extreme<br />

risk of extirpation across its range due to all seven of the<br />

following conditions: limited geographic range, small<br />

number of populations, low number of individuals, high<br />

habitat specificity, high intrinsic rarity, high threats, and<br />

downward population trend.<br />

High (6 points): species at high risk of extirpation<br />

rangewide or in the state. High priority species are<br />

scored as vulnerable for at least six of the seven ranking<br />

criteria.<br />

Watch (5 points): species currently secure but vulnerable<br />

to downward changes in status. These taxa are<br />

scored as vulnerable for at least five of the seven ranking<br />

criteria.<br />

Medium (4 points): species secure rangewide but<br />

vulnerable to extirpation in the state. Medium priority<br />

species are scored as vulnerable for at least four of the<br />

seven ranking criteria.<br />

Low (0-3 points): species secure rangewide and in<br />

the state. These species are scored as vulnerable for<br />

three or less of the seven ranking criteria.<br />

Need Data: insufficient data available to score species<br />

for at least three of the seven ranking criteria. If<br />

information were available. these species would likely<br />

be ranked as Extremely High, High, Watch, or Medium<br />

priority rather than Low priority.<br />

An example of ranking a <strong>Utah</strong> species<br />

The following example demonstrates the application<br />

of the Wyoming protocol. Penstemon gibbensii is a narrow<br />

endemic of extreme NE <strong>Utah</strong> (Daggett County),<br />

adjacent NW Colorado, and SC Wyoming, earning it 2<br />

points for geographic range. In <strong>Utah</strong>, it is known from a<br />

single occurrence in the Browns Park area (1 point for<br />

low number of populations) containing approximately<br />

700 plants (1 point for low number of individuals) (<strong>Utah</strong><br />

Division of Wildlife Resources 1998). It is restricted to<br />

barren white shales of the Browns Park Formation (1<br />

point for high habitat specificity). Little is known about<br />

the pollination biology or life history of P. gibbensii<br />

(Heidel 2009), suggesting an “unknown” score is appropriate<br />

for intrinsic rarity. Threats from trampling, soil<br />

erosion, and over-collection by gardeners are high<br />

throughout its range (1 point for threats). Trends in<br />

<strong>Utah</strong> are unknown, although some populations in Wyoming<br />

appear to be declining (Heidel 2009). The mini-<br />

199


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

mum score for P. gibbensii is 6 points, while the potential<br />

score is 8. The average of the two scores is 7, earning<br />

P. gibbensii a place on the UNPS Extremely High<br />

priority list.<br />

RESULTS<br />

The UNPS Rare <strong>Plant</strong> Committee met in January<br />

2009 to apply the Wyoming protocol to the entire flora<br />

of <strong>Utah</strong>. Based on the fourth edition of A <strong>Utah</strong> Flora<br />

(Welsh et al. 2008) and other recent literature (such as<br />

The Flora of North America and Intermountain Flora)<br />

we started with a pool of 4273 taxa* of vascular plants<br />

known to be native or introduced in <strong>Utah</strong>. We immediately<br />

removed 1017 cultivated and naturalized (nonnative)<br />

taxa, as we deemed these not to be of conservation<br />

importance in the state. Of the 3160 species native<br />

to <strong>Utah</strong>, we eliminated another 1421 common and widespread<br />

taxa (mostly ranked S4 or S5 by NatureServe) of<br />

*Taxa include full species and unique subspecies and varieties,<br />

treated here in the phylogenetic sense of Cracraft (1987)<br />

as the smallest evolutionary units that are diagnosably distinct.<br />

Table 2. Sample Ranking Form<br />

Species<br />

Date Scored<br />

Ranking Factors<br />

Evaluators<br />

Scores<br />

Select one score per ranking factor in either column A or B<br />

Column A<br />

1. Geographic Range Local Endemic (2) ____<br />

Regional Endemic,<br />

Disjunct, Peripheral,<br />

or Sparse (1) ____<br />

Widespread (0) ____<br />

2. <strong>Number</strong> of Populations Low (1) ____<br />

Medium to High (0) ____<br />

3. Abundance Low (1) ____<br />

Medium to High (0) ____<br />

4. Habitat Specificity High (1) ____<br />

Medium to Low (0) ____<br />

5. Intrinsic Rarity High (1) ____<br />

Medium to Low (0) ____<br />

6. Magnitude and Imminence<br />

of Threats<br />

High (1)<br />

Medium to Low (0)<br />

____<br />

____<br />

7. Population Trend Downward (1) ____<br />

Stable, Oscillating<br />

or Upward (0) ____<br />

TOTALS<br />

Conservation Priority*<br />

_____________________<br />

Sum of scores in<br />

Column A<br />

____<br />

Minimum<br />

(based on total score in<br />

Column A) ____<br />

Unknown (1)<br />

Unknown (1)<br />

Unknown (1)<br />

Unknown (1)<br />

Unknown (1)<br />

Unknown (1)<br />

Unknown (1)<br />

Column B<br />

Sum of scores in<br />

Column B<br />

____<br />

____<br />

____<br />

____<br />

____<br />

____<br />

____<br />

____<br />

Potential<br />

(based on sum of scores<br />

in Columns A + B) ____<br />

Comments<br />

Sum of scores in<br />

Column A + B<br />

____<br />

Averaged<br />

(based on average of<br />

scores in Columns<br />

A + B rounded down) ____<br />

*Conservation Priority is based on the averaged point total: Extremely high priority = total score of 7 or 8 points,<br />

High priority = total score of 6 points, Watch list = total score of 5, Medium priority = total score of 4 points, Low<br />

priority = total score of 0-3 points.<br />

200


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

low conservation priority. The remaining 1739 taxa<br />

were then ranked by the committee using the Wyoming<br />

protocol. At this point another 927 taxa were assigned<br />

to the Low Priority list, leaving 812 species among the<br />

other categories (Extremely High, High, Watch, Medium,<br />

and Need Data). These remaining taxa comprised<br />

the first draft of the UNPS rare plant list, presented at<br />

the 2009 Southwest Rare <strong>Plant</strong> Conference.<br />

Revisions suggested by the conference attendees and<br />

other specialists resulted in changes to the final score for<br />

over 100 taxa. These changes were reflected in the final<br />

UNPS rare plant list, published in the <strong>Society</strong>’s membership<br />

publication, the Sego Lily, in November 2009<br />

(Fertig 2009). The list has been revised twice since<br />

then. Twenty taxa were changed in rank due to new<br />

data and 11 newly described or discovered species were<br />

added (Fertig 2010, <strong>2012</strong>).<br />

Extremely High Priority At present, 31 <strong>Utah</strong> plant<br />

taxa are recognized as species of Extremely High conservation<br />

priority (Table 3, Appendix 1). This group<br />

represents just 1% of the entire flora of the state. To<br />

qualify as Extremely High priority, a species must have<br />

a limited geographic range, few populations, low number<br />

of individuals, high habitat specificity and intrinsic<br />

rarity, high threats, and a downward population trend.<br />

Nearly half (15 taxa) of these species are presently listed<br />

as Threatened or Endangered under the ESA, and another<br />

four species are candidates or proposed for listing.<br />

Another 14 of these species are designated as Sensitive<br />

by the BLM or Forest Service. Just two of the 31 Extremely<br />

High priority plant taxa lack any formal designation:<br />

Iris pariensis (a species with taxonomic questions<br />

that may be extinct) and Viola clauseniana<br />

(endemic to Zion National Park and increasingly threatened<br />

by competition from exotic plants and possible<br />

over-collection; Fertig 2010b). Since the initial UNPS<br />

list was published one species has been dropped from<br />

the Extremely High priority list (Sclerocactus wetlandicus,<br />

changed to High priority) and one species has<br />

been added (Carex specuicola, upgraded from the High<br />

priority list).<br />

High Priority The High priority list presently contains<br />

119 taxa, up from 114 recognized in 2009 (Table<br />

3, Appendix 2). This group consists of only 3.8% of the<br />

native flora of <strong>Utah</strong>. High priority plant species generally<br />

have limited geographic ranges, low population<br />

size, few known occurrences, and high habitat specificity,<br />

but have lower intrinsic rarity, fewer threats, or stable<br />

trends (or these factors are unknown) compared to<br />

Extremely High priority taxa. Eight of the state’s 24<br />

federally listed Threatened or Endangered species are<br />

ranked High priority by UNPS, and two other species<br />

from this group are currently candidates for potential<br />

listing. Half of the High priority taxa (60 species) are<br />

listed as Sensitive by the BLM or Forest Service. The<br />

remaining 51 species have no status and consist mostly<br />

of species that were once considered for listing under<br />

the ESA (41 former Category 2 or 3C taxa), or are recently<br />

described (Atwood et al, 1991; Welsh and Atwood<br />

2009).<br />

Watch List The UNPS Watch list currently contains<br />

264 taxa (Table 3, Appendix 3), or 8.4% of the native<br />

<strong>Utah</strong> flora. This category contains local or regional<br />

endemics with high habitat specificity or intrinsic rarity,<br />

but which are either locally abundant or apparently secure<br />

at present. If current conditions were to change<br />

significantly, however, these species could easily trend<br />

downward and become species of greater concern (and<br />

be rescored as Extremely High or High priority). At<br />

least 115 species in the Watch category were initially<br />

considered for potential listing under the ESA (Atwood<br />

et al. 1991, Ayensu and DeFilipps 1978; Greenwalt<br />

1975, Welsh 1978, Welsh et al. 1975, Welsh and Chatterley<br />

1985) in the 1970s and 1980s, but were subsequently<br />

dropped from consideration after better survey<br />

data found them to be less imminently threatened or<br />

rare. Astragalus montii is the only species currently<br />

listed under the ESA that is categorized on the Watch<br />

list (Erigeron maguirei, recently delisted, is also on the<br />

Watch list). Another 70 species in the Watch category<br />

are presently listed as Sensitive by the BLM or Forest<br />

Service. Over one-third of the changes to the UNPS list<br />

since 2009 have involved additions or status changes<br />

affecting the Watch category.<br />

Medium Priority Another 329 plant species in <strong>Utah</strong><br />

are currently ranked as Medium priority for conservation<br />

attention (Table 3)*. Species in this category are<br />

usually widespread globally but rare within the state,<br />

with few known occurrences, low numbers, or potentially<br />

high threats. A small subset of Medium priority<br />

species (including 13 listed as Sensitive by the BLM or<br />

Forest Service) are locally common regional endemics<br />

with stable trends and low threats that might otherwise<br />

be treated on the Watch list. Medium priority taxa account<br />

for 10.4% of the native flora of <strong>Utah</strong>.<br />

Need Data A total of 115 taxa are presently on the<br />

UNPS Need Data list (Table 3, Appendix 4), or 3.6% of<br />

the state’s native flora. This is the fastest growing category,<br />

as it tends to be the repository for newly described<br />

species or those discovered for the first time within<br />

<strong>Utah</strong>. Recent additions include Astragalus kelseyae,<br />

Eremogone loisiae, Eriogonum domitum, and Navarretia<br />

furnissii, all named as new species since 2009<br />

(Corbin 2011, Grady and Reveal 2011; Holmgren and<br />

Holmgren 2011, Johnson et al. <strong>2012</strong>). The Need Data<br />

list also includes species with unresolved taxonomic<br />

questions and those that have been reported for the state<br />

*In the interests of space, the Medium priority list is not included<br />

in this paper, but is available by request from UNPS.<br />

201


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Figure 1. <strong>Utah</strong> Counties with three-letter codes used in<br />

Table 3 and Appendix 1-4.<br />

Figure 2. TNC Ecoregions of <strong>Utah</strong>, used in Table 3, as<br />

defined by Stein et al. (2000).<br />

Table 3. Summary of UNPS Rare <strong>Plant</strong> List, 2009-<strong>2012</strong><br />

See text for explanation and scoring of each of the ranking categories. Counties are depicted in Figure 1. Ecoregions<br />

are defined as geographic regions with similar climate, topography, and vegetation, as defined by The Nature Conservancy<br />

(Stein et al. 2000) and are shown in Figure 2.<br />

State/County/TNC Ecoregion<br />

Extremely<br />

High<br />

202<br />

High Watch Need<br />

Data<br />

Medium<br />

State <strong>Utah</strong> Statewide 31 119 264 115 329 858<br />

County<br />

Beaver County (Bvr) 0 8 13 8 24 53<br />

Box Elder County (Box) 1 2 10 6 28 47<br />

Cache County (Cch) 0 3 7 4 28 42<br />

Carbon County (Crb) 1 1 9 5 3 19<br />

Daggett County (Dag) 1 2 13 3 16 35<br />

Davis County (Dav) 0 0 2 2 6 10<br />

Duchesne County (Dch) 4 14 28 9 24 79<br />

Emery County (Emr) 5 8 24 17 23 77<br />

Garfield County (Grf) 1 17 49 17 49 133<br />

Grand County (Grn) 1 12 26 17 29 85<br />

Iron County (Irn) 0 5 18 7 23 53<br />

Juab County (Jub) 1 6 13 12 17 49<br />

Kane County (Kan) 3 25 52 12 67 159<br />

Millard County (Mil) 0 6 21 17 21 65<br />

Total


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Table 3. Continued<br />

County<br />

TNC<br />

Ecoregion<br />

State/County/TNC Ecoregion<br />

Extremely<br />

High<br />

High Watch Need<br />

Data<br />

Medium<br />

Morgan County (Mor) 0 0 2 0 1 3<br />

Piute County (Piu) 0 9 13 3 14 39<br />

Rich County (Rch) 0 1 4 5 9 19<br />

Salt Lake County (Slt) 0 8 12 2 20 42<br />

San Juan County (Snj) 2 13 39 15 63 132<br />

Sanpete County (Snp) 2 7 12 4 8 33<br />

Sevier County (Sev) 3 13 18 6 21 61<br />

Summit County (Sum) 0 1 5 3 15 24<br />

Tooele County (Toe) 1 3 13 5 13 35<br />

Uintah County (Uin) 6 15 36 10 23 90<br />

<strong>Utah</strong> County (Uta) 1 9 14 12 14 50<br />

Wasatch County (Was) 0 3 8 4 8 23<br />

Washington County (Wsh) 6 18 76 17 115 232<br />

Wayne County (Way) 6 12 13 11 22 64<br />

Weber County (Web) 0 3 6 4 9 22<br />

Colorado Plateau (CP) 13 47 98 41 124 323<br />

Columbia River Basin (CRB) 1 1 2 2 19 25<br />

Great Basin (GB) 2 16 43 28 63 152<br />

Mohave Desert (MD) 4 11 44 12 67 138<br />

<strong>Utah</strong> High Plateaus (UHP) 9 36 69 28 63 205<br />

<strong>Utah</strong>-Wyoming Rocky Mtns<br />

(UWRM)<br />

Total<br />

0 17 50 16 65 148<br />

Wyoming Basins (WyB) 7 12 21 11 19 70<br />

in the literature with ambiguous supporting data. The<br />

majority of taxa in this category need additional information<br />

on trend, threats, and abundance. At least 15 of<br />

these species are designated as BLM or Forest Service<br />

Sensitive. Eriogonum corymbosum var. nilesii is a candidate<br />

for potential listing under the ESA, although a<br />

recent monograph questions the veracity of <strong>Utah</strong> reports<br />

(Reveal in Holmgren et al. <strong>2012</strong>).<br />

Low Priority The remaining 2302 native plant taxa of<br />

<strong>Utah</strong> are presently scored as Low conservation priority<br />

(72.8% of the total native flora). These species are usually<br />

widespread rangewide and within <strong>Utah</strong>, have numerous<br />

populations in the state, low habitat specificity<br />

and intrinsic rarity, stable to increasing trends, and few<br />

threats. These species are still important for the sake of<br />

conserving overall biodiversity, but rarely warrant individualized<br />

attention.<br />

All told, 858 of <strong>Utah</strong>’s 3160 native plant taxa (27.2%)<br />

have been identified as Extremely High, High, Watch,<br />

or Medium priority, or are on the UNPS need data list<br />

using the Wyoming protocol system. The distribution<br />

of these species across the state is not random. With<br />

232 taxa of conservation concern (27% of the state total),<br />

Washington County has the highest number of species<br />

of conservation concern of any county in <strong>Utah</strong><br />

(Figure 1, Table 3). This high richness can be explained<br />

in part by Washington County’s location at the confluence<br />

of four major floristic regions: the Mojave Desert,<br />

203


Great Basin, Colorado Plateau, and Rocky Mountains.<br />

The next three counties with the greatest number of<br />

species of conservation concern (Kane, Garfield, and<br />

San Juan) are all, like Washington County, located on or<br />

near the southern boundary of the state. Additional<br />

counties with a high number of species of concern include<br />

Uintah, Duchesne, Grand, and Emery counties<br />

near the eastern border of <strong>Utah</strong>. By contrast, the number<br />

of rare species is relatively low in the northern and<br />

western tier of counties. Surprisingly few plant species<br />

of conservation concern occur in the greater Salt Lake<br />

City area, though this may be an artifact of undersampling<br />

or reflect significant habitat losses over the<br />

last 150 years of settlement (Fertig 2009).<br />

Ecoregions are defined as geographic areas with a<br />

similar climate, topography, and vegetation. The Nature<br />

Conservancy has developed a national ecoregional classification<br />

(Stein et al. 2000) that recognizes seven ecoregions<br />

in <strong>Utah</strong> (Figure 2*). Of these, the Colorado Plateau<br />

ecoregion has the highest number of plant species<br />

of conservation concern with 323 taxa, or 37.6% of the<br />

state total (Table 3). This region, which includes the<br />

canyon country and La Sal and Abajo mountains of<br />

southeast <strong>Utah</strong>, also has the highest number of endemic<br />

species in the state (Welsh and Atwood 2009). Although<br />

comparable in area to the Colorado Plateau, the<br />

Great Basin ecoregion has less than half as many species<br />

of concern (152 taxa). The Mohave Desert ecoregion<br />

of extreme southwestern <strong>Utah</strong> is the second smallest<br />

in area in the state (after the Columbia River Basin<br />

in the Grouse Creek and Raft River mountains of northwest<br />

<strong>Utah</strong>) but has the highest concentration of species<br />

of concern per unit area (138 taxa in all). The <strong>Utah</strong><br />

High Plateaus, which extends from the Tavaputs Plateau<br />

and Book Cliffs of eastern <strong>Utah</strong> to the Wasatch Plateau,<br />

and Markagunt and Paunsaugunt plateaus of southcentral<br />

<strong>Utah</strong>, has the second highest concentration of<br />

endemics and taxa of conservation concern (205 species,<br />

or 23.9% of the state total) (Table 3).<br />

DISCUSSION<br />

The UNPS rare plant list is just the latest in a long<br />

series of comparable publications dating back to the<br />

passage of the Endangered Species Act (ESA) of 1973.<br />

No plants were included in the very first official list of<br />

species protected under the ESA, but Congress directed<br />

the Smithsonian Institution to develop the first national<br />

list of vascular plants that might qualify for listing as<br />

*Welsh and Atwood (2009) have developed a similar system<br />

of “geoendemic areas” to identify floristic regions of <strong>Utah</strong>.<br />

Their map depicts 12 subregions and differs from the TNC<br />

system in lumping the Columbia River Basin with the Great<br />

Basin and in more finely subdividing the <strong>Utah</strong>-Wyoming<br />

Rocky Mountains, <strong>Utah</strong> High Plateaus, Colorado Plateau, and<br />

Mohave Desert ecoregions.<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

204<br />

Threatened or Endangered. That list appeared in 1975<br />

and was based on the best available information at the<br />

time (Ayensu and DeFilipps 1978, Greenwalt 1975).<br />

The Smithsonian Institution cited 761 plant taxa as potentially<br />

Endangered, 1238 as Threatened, and 100 as<br />

extinct in the continental United States (another 1088<br />

endangered, threatened, and extinct species were reported<br />

for Hawaii). Of these, 156 species were from<br />

<strong>Utah</strong>, including 56 listed as endangered, 91 threatened,<br />

and 9 extinct (Greenwalt 1975).<br />

Welsh and others (1975) reviewed the Smithsonian<br />

publication and developed the first <strong>Utah</strong>-specific compilation<br />

of endangered, threatened, extinct, endemic, and<br />

rare plant species in 1975. Welsh and his co-authors<br />

recognized 66 <strong>Utah</strong> plant taxa as possibly endangered,<br />

198 as threatened, 7 as extinct, and 20 as extirpated<br />

(extinct in <strong>Utah</strong>, but extant elsewhere). Most of the recommendations<br />

by Welsh and others (1975) were incorporated<br />

into a revised Smithsonian list (Ayensu and De-<br />

Filipps 1978) that became part of a proposal to list<br />

nearly 1700 plant species as Threatened or Endangered<br />

in 1976 (the proposal was ultimately dismissed).<br />

These initial rare species lists were plagued by incomplete<br />

data and taxonomic problems. Of the 156<br />

<strong>Utah</strong> species considered endangered, threatened, or extinct<br />

by the Smithsonian Institution in 1975, only 39<br />

(25%) are still considered taxa of Extremely High or<br />

High conservation priority today. At least 13 of these<br />

species (8.3%) are no longer recognized as legitimate<br />

taxa. Another 28 species (18%) are now known to be<br />

much more common or less threatened and are classified<br />

as Low priority by UNPS. Eight of the nine species<br />

considered extinct in 1975 have been rediscovered (only<br />

Cuscuta warneri is still thought to be extirpated in<br />

<strong>Utah</strong>). Among the additional 225 state endemics and<br />

other potentially rare species evaluated by Welsh and<br />

others (1975), one half (112 taxa) are now scored as<br />

Low priority and 26 (11.5%) are no longer recognized<br />

taxonomically.<br />

Over the next two decades new <strong>Utah</strong> rare plant lists<br />

were developed by the US Fish and Wildlife Service,<br />

Bureau of Land Management, US Forest Service, and<br />

non-governmental organizations (such as The Nature<br />

Conservancy and <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>). The composition<br />

of these lists continued to evolve to reflect<br />

ever-improving knowledge of the distribution, abundance,<br />

and status of the state’s flora (Atwood et al.<br />

1991; <strong>Utah</strong> Division of Wildlife Resources 1998; <strong>Utah</strong><br />

<strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> 1980, 1982; Welsh 1978; Welsh<br />

and Chatterley 1985; Welsh and Thorne 1979). Threat<br />

of potential listings under the ESA prompted a large<br />

scale effort to survey rare species and remote corners of<br />

the nation for new taxa. During the period from 1975 to<br />

1994 nearly 1200 new vascular plant taxa were described<br />

across North America, or approximately 60 new


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

species per year (Hartman and Nelson 1998). In <strong>Utah</strong><br />

alone, over 250 new plant taxa have been named since<br />

the early 1970s (Welsh et al. 2008). Not surprisingly,<br />

most of these newly discovered species are narrow endemics<br />

with few or small populations and specialized<br />

habitat requirements, making them potential candidates<br />

for rare plant lists. Fifty-five percent of the current<br />

UNPS Extremely High priority list (17 species) and<br />

52% of the High priority list (62 species) have only<br />

been named since 1978.<br />

In the years since 1975, field surveys and taxonomic<br />

research have resulted in many plant taxa being removed<br />

from consideration as species of concern due to<br />

lack of threat, stable trends, or documented abundance.<br />

At least 772 of the current 2302 <strong>Utah</strong> plant taxa scored<br />

as Low priority by UNPS (33.5%) have been listed as<br />

potentially Endangered, Threatened, extinct, or otherwise<br />

rare at one time. Of the 28 <strong>Utah</strong> species that have<br />

been listed as Threatened or Endangered by the US Fish<br />

and Wildlife Service since 1978, four have subsequently<br />

been delisted (Astragalus perianus, Erigeron maguirei,<br />

Echinocereus engelmannii var. purpureus, and Echinocereus<br />

triglochidiatus var. inermis) because surveys<br />

have found them to be much more common, or the taxa<br />

are no longer recognized.<br />

Future Applications<br />

Rare plant lists have a short shelf life. The UNPS list<br />

has already been revised twice since it first appeared in<br />

2009 and will need to be updated again in the coming<br />

year. With the publication of the final volume of the<br />

Intermountain Flora (Holmgren et al. <strong>2012</strong>) at least 36<br />

new native plant species have been documented in <strong>Utah</strong><br />

which have not been evaluated by the UNPS Rare <strong>Plant</strong><br />

Committee. Several of these species are narrow endemics<br />

that are likely to be ranked as Extremely High, High,<br />

or Watch list species when sufficient data are available<br />

for review. Other species currently on the Need Data<br />

list will also likely be placed in higher priority categories<br />

in the near future. Undoubtedly, there are more rare<br />

species still awaiting discovery in the years ahead. Results<br />

of on-going monitoring studies and field inventories<br />

will also improve our understanding of many species<br />

and result in shifts in their conservation priority.<br />

In addition to <strong>Utah</strong>, the Wyoming protocol has been<br />

recently applied to the entire flora of Wyoming (Fertig<br />

2011) and to Zion National Park (Fertig 2010b). In<br />

Zion, the park’s initial list of over 200 species of concern<br />

was streamlined to 51 taxa, of which only 13 were<br />

deemed Extremely High or High priority. Some species<br />

were given a slightly different rank in the park compared<br />

to the state as a whole, reflecting differences in<br />

scale and data sufficiency (Fertig 2010b). The Idaho<br />

and Arizona native plant societies have also expressed<br />

interest in using this methodology to rank rare plants in<br />

their respective states. As it is used more frequently, the<br />

protocol will hopefully be strengthened and improved.<br />

It is important to remember that the UNPS rare plant<br />

list has no binding legal authority and is only as<br />

accurate as the information used for ranking. The list<br />

and the listing process remain useful, however, because<br />

they provide a simple, repeatable, and transparent<br />

method to prioritize conservation action among hundreds<br />

of rare species. With conservation resources<br />

stretched thin and time running out, this form of triage<br />

may be critical to preserving <strong>Utah</strong>’s most vulnerable<br />

botanical treasures.<br />

ACKNOWLEDGMENTS<br />

These lists were developed with the input of my fellow<br />

members of the <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> Rare<br />

<strong>Plant</strong> Committee: Duane Atwood (BYU herbarium, retured),<br />

Rita [Dodge] Reisor (Red Butte Garden), Robert<br />

Fitts (UT Conservation Data Center), Ben Franklin (UT<br />

Conservation Data Center, retired), and Jason Alexander<br />

(<strong>Utah</strong> Valley University). The original list benefited<br />

from input from more than 40 participants in the <strong>Utah</strong><br />

rare plant breakout session at the 2009 Southwest Rare<br />

<strong>Plant</strong> Conference. Additional helpful comments and<br />

suggestions have been provided by attendees of the annual<br />

<strong>Utah</strong> Rare <strong>Plant</strong> meetings from 2010 to <strong>2012</strong>. Special<br />

thanks to the following for their input on various<br />

iterations of the UNPS list: Ron Bolander (BLM state<br />

botanist), Jessie Brunson (USFWS), Debi Clark<br />

(Canyon De Chelley NM), Cheryl Decker (NPS SE<br />

<strong>Utah</strong> Group), Larry England (USFWS, retired), Tony<br />

Frates (UNPS Conservation Committee and webmaster),<br />

Kipp Lee (UNPS), Kezia Nielson (Zion NP),<br />

Teresa Prendusi (USFS regional botanist), Gary Reese<br />

(consultant), Daniela Roth (formerly USFWS), Jim<br />

Spencer (NRCS, Roosevelt UT), Blake Wellard<br />

(University of <strong>Utah</strong> grad student) and Dorde Woodruff<br />

(<strong>Utah</strong> cactus expert). My apologies (and thanks) to<br />

other contributors whom I have omitted inadvertently.<br />

LITERATURE CITED<br />

Akçakaya, H.R., S. Ferson, M.A. Burgman, D.A.<br />

Keith, G.M. Mace, and C.R. Todd. 2000. Making consistent<br />

IUCN classifications under uncertainty. Conservation<br />

Biology 14 (4):1001-1013.<br />

Andelman, S.J., C. Groves, and H.M. Regan. 2004.<br />

A review of protocols for selecting species at risk in the<br />

context of US Forest Service viability assessments.<br />

Acta Oecologica 26:75-83.<br />

Atwood, D., J. Holland. R. Bolander, B. Franklin,<br />

D.E. House, L. Armstrong, K. Thorne, and L. England.<br />

1991. <strong>Utah</strong> Threatened, Endangered, and Sensitive<br />

<strong>Plant</strong> Field Guide. US Forest Service Intermountain<br />

Region, National Park Service, Bureau of Land Management,<br />

<strong>Utah</strong> Natural Heritage Program, US Fish and<br />

205


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Wildlife Service, Environmental Protection Agency,<br />

Navajo Nation, and Skull Valley Goshute Tribe.<br />

Ayensu, E.S. and R.A. DeFilipps. 1978. Endangered<br />

and Threatened <strong>Plant</strong>s of the United States. Smithsonian<br />

Institution and World Wildlife Fund, Washington,<br />

DC. 403 pp.<br />

Beissinger, S.R., J.M. Reed, J.M. Wunderle Jr., S.K.<br />

Robinson, and D.M. Finch. 2000. Report of the AOU<br />

conservation committee on the Partners in Flight species<br />

prioritization plan. The Auk 117:549-561.<br />

Breininger, D.R., M. Barkaszi, R.B. Smith, D.M.<br />

Oddy, and J.A. Provancha. 1998. Prioritizing wildlife<br />

taxa for biological diversity conservation at the local<br />

scale. Environmental Management 22:315-321.<br />

Center for <strong>Plant</strong> Conservation. 2000. America’s<br />

Vanishing Flora, Stories of Endangered <strong>Plant</strong>s from the<br />

Fifty States and Efforts to Save Them. Center for <strong>Plant</strong><br />

Conservation, St. Louis, MO. 59 pp.<br />

Corbin, B.L. 2011. A new species of Astragalus<br />

(Fabaceae) from the Wasatch Mountains of <strong>Utah</strong>.<br />

Madroño 58(3):185-189.<br />

Cracraft, J. 1987. Species concepts and the ontology<br />

of evolution. Biology and Philosophy 2:329-346.<br />

Faber-Langendoen, D., L. Master, J. Nichols, K.<br />

Snow, A. Tomaino, R. Bittman, G. Hammerson, B. Heidel,<br />

L. Ramsay, and B. Young. 2009. NatureServe conservation<br />

status assessments: Methodology for assigning<br />

ranks. NatureServe, Arlington, VA. 42 pp.<br />

Fertig, W. 2009. Developing a <strong>Utah</strong> rare plant list.<br />

Sego Lily 32(6):1-17.<br />

Fertig, W. 2010a. Updates to UNPS rare plant list.<br />

Sego Lily 33(6):15.<br />

Fertig, W. 2010b. Rare plants of Zion National<br />

Park. Moenave Botanical Consulting, Kanab, UT. 95<br />

pp.<br />

Fertig, W. 2011. Strategies for plant conservation in<br />

Wyoming: Distributional modeling, gap analysis, and<br />

identifying species at risk. Doctoral dissertation, Department<br />

of Botany, University of Wyoming, Laramie,<br />

WY. 451 pp.<br />

Fertig, W. <strong>2012</strong>. Rare plant committee updates<br />

UNPS rare plant list. Sego Lily 35(3):7.<br />

Grady, B.R. and J.L. Reveal. 2011. New combinations<br />

and a new species of Eriogonum (Polygonaceae:<br />

Eriogonoideae) from the Great Basin Desert, United<br />

States. Phytotaxa 24:33-38.<br />

Greenwalt, L.A. 1975. Endangered and threatened<br />

wildlife and plants. Federal Register 40:44412-44429.<br />

Hartman, R.L. and B.E. Nelson. 1998. Taxonomic<br />

novelties from North America north of Mexico: A 20-<br />

year vascular plant diversity baseline. Monographs in<br />

Systematic Botany from the Missouri Botanical Garden<br />

67:1-59.<br />

Heidel, B. 2009. Survey and monitoring of Gibbens’<br />

penstemon (Penstemon gibbensii) in south-central<br />

Wyoming. Report prepared for the Bureau of Land<br />

Management. Wyoming Natural Diversity Database,<br />

Laramie, WY. 36 pp.<br />

Holmgren, N.H. and P.K. Holmgren. 2011. A new<br />

species of Eremogone (Caryophyllaceae) from northern<br />

<strong>Utah</strong> and southeastern Idaho, U.S.A. Brittonia 63(1):1-<br />

6.<br />

Holmgren, N.H., P.K. Holmgren, J.L. Reveal, and<br />

collaborators. <strong>2012</strong>. Intermountain Flora, Vascular<br />

<strong>Plant</strong>s of the Intermountain U.S.A. Volume two, part A,<br />

subclasses Magnoliidae-Caryophyllidae. New York Botanical<br />

Garden, Bronx, NY. 731 pp.<br />

Holsinger, K.E. 1992. Setting priorities for regional<br />

plant conservation programs. Rhodora 94 (879):243-<br />

257.<br />

IUCN. 2001. IUCN Red List Categories and Criteria<br />

Version 3.1. IUCN Species Survival Commission.<br />

IUCN, Gland, Switzerland and Cambridge, UK. 30 pp.<br />

Johnson, L.A., L.M. Chan, K. Burr, and D. Hendrickson.<br />

<strong>2012</strong>. Navarretia furnissii (Polemoniaceae),<br />

a new diploid species from the intermountain western<br />

United States distinguished from tetraploid Navarretia<br />

saximontana. Phytotaxa 42:51-61.<br />

Keith, D.A. 1998. An evaluation and modification<br />

of World Conservation Union red list criteria for classification<br />

of extinction risk in vascular plants. Conservation<br />

Biology 12 (5):1076-1090.<br />

Mace, G.M., N.J. Collar, K.J. Gaston, C. Hilton-<br />

Taylor, H.R. Akçakaya, N. Leader-Williams, E.J.<br />

Milner-Gulland, and S.N. Stuart. 2008. Quantification<br />

of extinction risk: IUCN’s system for classifying threatened<br />

species. Conservation Biology 22(6):1424-1442.<br />

Master, L.L. 1991. Assessing threats and setting<br />

priorities for conservation. Conservation Biology 5:<br />

148-157.<br />

Master, L.L., B.A. Stein, L.S. Kutner, and G.A.<br />

Hammerson. 2000. Vanishing assets, conservation<br />

status of U.S. species. Pp. 93-118. In: Stein, B.A., L.S.<br />

Kutner, and J.S. Adams, eds. Precious Heritage, the<br />

Status of Biodiversity in the United States. The Nature<br />

Conservancy and Association for Biodiversity Information,<br />

Oxford University Press, New York.<br />

Millsap, B.A., J.A. Gore, D.E. Runde, and S.L. Cerulean.<br />

1990. Setting priorities for the conservation of<br />

fish and wildlife species in Florida. Wildlife Monographs<br />

111:1-57.<br />

Noss, R.F. and A.Y. Cooperrider. 1994. Saving Nature’s<br />

Legacy, Protecting and Restoring Biodiversity.<br />

Island Press, Washington, DC. 416 pp.<br />

O’Grady, J.J., M.A. Burgman, D.A. Keith, L.L. Master,<br />

S.J. Andelman, B.W. Brook, G.A. Hammerson, T.<br />

Regan, and R. Franklin. 2004. Correlations among extinction<br />

risks assessed by different systems of threatened<br />

species categorization. Conservation Biology 18<br />

(6):1624-1635.<br />

206


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Panjabi, A.O., E.H. Dunn, P.J. Blancher, W.C.<br />

Hunter, B. Altman, J. Bart, C.J. Beardmore, H. Berlanga,<br />

G.S. Butcher, S.K. Davis, D.W. Demarest, R.<br />

Dettmers, W. Easton, H. Gomez de Silva Garza, E.E.<br />

Iñigo-Elias, D.N. Pashley, C.J. Ralph, T.D. Rich, K.V.<br />

Rosenberg, C.M. Rustay, J.M. Ruth, J.S. Wendt, and<br />

T.C. Will. 2005. The Partners in Flight handbook on<br />

species assessment. Version 2005. Partners in Flight<br />

Technical Series No. 3. Rocky Mountain Bird Observatory<br />

website: http://www.rmbo.org/downloads/<br />

Handbook2005.pdf.<br />

Rabinowitz, D. 1981. Seven forms of rarity. Pp.<br />

205-217. In: Synge, H. (ed.). The Biological Aspects of<br />

Rare <strong>Plant</strong> Conservation. John Wiley and Sons, Ltd.<br />

Regan, T.J., M.A. Burgman, M.A. McCarthy, L.L.<br />

Master, D.A. Keith, G.M. Mace, and S.J. Andelman.<br />

2005. The consistency of extinction risk classification<br />

protocols. Conservation Biology 19 (6):1969-1977.<br />

Regan, T.J., L.L. Master, and G.A. Hammerson.<br />

2004. Capturing expert knowledge for threatened species<br />

assessments: a case study using NatureServe conservation<br />

status rank. Acta Oecologica 26:95-107.<br />

Spence, J. <strong>2012</strong>. A new look at ranking plant rarity<br />

for conservation purposes, with an emphasis on the flora<br />

of the American Southwest. Calochortiana 1:25-34.<br />

Stein, B.A., L.S. Kutner, and J.S. Adams. 2000.<br />

Precious Heritage, the Status of Biodiversity in the<br />

United States. The Nature Conservancy and Association<br />

for Biodiversity Information, Oxford University<br />

Press, New York. 399 pp.<br />

US Fish and Wildlife Service. 1983. Endangered<br />

and Threatened species listing and recovery priority<br />

guidelines. Federal Register 48 (184):43098-43105.<br />

<strong>Utah</strong> Division of Wildlife Resources. 1998. Inventory<br />

of sensitive species and ecosystems in <strong>Utah</strong>. Endemic<br />

and rare plants of <strong>Utah</strong>: an overview of their distribution<br />

and status. Report prepared for <strong>Utah</strong> Reclamation<br />

Mitigation and Conservation Commission and US<br />

Department of Interior. 566 pp. + app.<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 1980. <strong>Utah</strong> threatened<br />

and endangered plants. <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Newsletter 3(1):1-4.<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong>. 1982. Report on <strong>Utah</strong><br />

rare/threatened/endangered plant conference. Sego Lily<br />

5(1):5-9.<br />

Welsh, S.L. 1978. Endangered and threatened plants<br />

of <strong>Utah</strong>: a reevaluation. Great Basin Naturalist 38:1-18.<br />

Welsh, S.L. and N.D. Atwood. 2009. <strong>Plant</strong> Endemism<br />

and Geoendemic Areas of <strong>Utah</strong>. Privately published.<br />

97 pp.<br />

Welsh, S.L., N.D. Atwood, S. Goodrich, and L.C.<br />

Higgins. 2008. A <strong>Utah</strong> Flora, 2004-2008 summary<br />

monograph, fourth edition, revised. Brigham Young<br />

University, Provo, UT. 1019 pp.<br />

Welsh, S.L., N.D. Atwood, and J.L. Reveal. 1975.<br />

Endangered, threatened, extinct, endemic, and rare or<br />

restricted <strong>Utah</strong> vascular plants. Great Basin Naturalist<br />

35:327-376.<br />

Welsh, S.L. and L.M. Chatterley. 1985. <strong>Utah</strong>’s rare<br />

plants revisited. Great Basin Naturalist 45:173-236.<br />

Welsh, S.L. and K.H. Thorne. 1979. Illustrated<br />

Manual of Proposed Endangered and Threatened <strong>Plant</strong>s<br />

of <strong>Utah</strong>. US Fish and Wildlife Service, Bureau of Land<br />

Management, and US Forest Service. 316 pp.<br />

207


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 1. UNPS Rare <strong>Plant</strong> List: Extremely High Priority Species<br />

The following table lists 31 species scored as Extremely High priority for conservation attention in <strong>Utah</strong> based on the<br />

Wyoming protocol ranking system. Species are listed alphabetically by family and scientific name. Synonyms for<br />

family and species names are included in parentheses. See text for an explanation of the seven ranking criteria and<br />

scoring methods used to derive the minimum and potential scores. County codes are explained in Table 3. Legal<br />

Status: Bureau of Land Management (BLM) and US Forest Service (USFS) Sensitive = S; US Fish and Wildlife Service<br />

(USFWS) Candidate = C, Endangered = E, Proposed = P; Threatened = T.<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Agavaceae<br />

Yucca sterilis<br />

(Y. harrimaniae var. s.)<br />

Creeping yucca 2 1 1 U 1 1 U 6 8 Dch?, Uin;<br />

BLM:S<br />

Asteraceae<br />

(Compositae)<br />

Townsendia aprica<br />

Last Chance townsendia<br />

2 1 1 1 0 1 1 7 7 Emr, Sev, Way;<br />

USFWS: T<br />

Brassicaceae<br />

(Cruciferae)<br />

Lepidium barnebyanum<br />

Schoenocrambe<br />

argillacea<br />

(Hesperidanthus a.)<br />

Schoenocrambe<br />

barnebyi<br />

(Hesperidanthus b.)<br />

Schoenocrambe<br />

suffrutescens<br />

(Hesperidanthus s.)<br />

Barneby’s pepperwort<br />

2 1 1 1 U 1 U 6 8 Dch; USFWS: E<br />

Clay reed-mustard 2 1 1 1 U 1 U 6 8 Uin; USFWS: T<br />

Barneby’s reedmustard<br />

Shrubby reedmustard<br />

Cactaceae Pediocactus despainii Despain’s pincushion<br />

cactus<br />

2 1 1 1 U 1 U 6 8 Emr, Way;<br />

USFWS: E<br />

2 1 1 1 U 1 1 7 8 Dch, Uin;<br />

USFWS: E<br />

2 1 1 1 1 1 1 8 8 Emr, Way;<br />

USFWS: E<br />

Pediocactus winkleri<br />

Sclerocactus brevispinus<br />

(S. whipplei var. ilseae)<br />

Sclerocactus wrightiae<br />

Winkler’s pincushion<br />

cactus<br />

Pariette hookless<br />

cactus<br />

Wright’s fishhook<br />

cactus<br />

2 1 1 1 1 1 1 8 8 Way; USFWS: T<br />

2 1 1 1 1 1 U 7 8 Dch, Uin;<br />

USFWS: T<br />

2 0 1 1 1 1 1 7 7 Emr, Way;<br />

USFWS: E<br />

Chenopodiaceae Atriplex canescens var.<br />

gigantea<br />

(Not recognized by<br />

Holmgren et al. <strong>2012</strong>)<br />

Giant fourwing<br />

saltbush<br />

2 1 1 1 U 1 U 6 8 Jub; BLM: S<br />

Cyperaceae Carex specuicola Navajo sedge 1 1 1 1 1 1 1 7 7 Snj; USFWS:T<br />

Originally on<br />

High priority list<br />

Fabaceae<br />

(Leguminosae)<br />

Astragalus<br />

ampullarioides<br />

Shivwits milkvetch<br />

2 1 1 1 1 1 1 8 8 Wsh; USFWS: E<br />

Astragalus anserinus<br />

Goose Creek milkvetch<br />

2 1 1 1 0 1 1 7 7 Box; BLM: S;<br />

USFS: S;<br />

USFWS: C<br />

208


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 1. UNPS Rare <strong>Plant</strong> List: Extremely High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Fabaceae<br />

(Leguminosae)<br />

Astragalus holmgreniorum<br />

Holmgren’s milkvetch<br />

2 1 1 1 1 1 1 8 8 Wsh; USFWS: E<br />

Astragalus iselyi Isely’s milkvetch 2 1 1 1 0 1 1 7 7 Grn, Snj;<br />

BLM:S; USFS:S<br />

Astragalus lentiginosus<br />

var. pohlii<br />

Pohl’s milkvetch 2 1 1 1 1 1 U 7 8 Toe; BLM: S<br />

Trifolium variegatum<br />

var. parunuweapensis<br />

Parunuweap<br />

clover<br />

2 1 1 1 U 1 U 6 8 Kan; BLM: S<br />

Hydrophyllaceae Phacelia argillacea Clay phacelia 2 1 1 1 0 1 1 7 7 <strong>Utah</strong>; USFWS: E<br />

Iridaceae<br />

Lamiaceae<br />

(Labiatae)<br />

Phacelia utahensis <strong>Utah</strong> phacelia 2 1 1 1 U 1 1 7 8 Snp, Sev;<br />

BLM:S<br />

Iris pariensis<br />

(Included in Iris missouriensis<br />

in FNA)<br />

Salvia columbariae var.<br />

argillacea<br />

Paria iris 2 1 1 U U 1 1 6 8 Kan<br />

Chinle chia 2 1 1 1 1 1 1 8 8 Kan, Wsh; BLM:<br />

S<br />

Loasaceae Mentzelia argillosa Arapien stickleaf 2 1 1 1 0 1 1 7 7 Snp, Sev; BLM:<br />

S<br />

Malvaceae Sphaeralcea gierischii Gierisch’s globemallow<br />

2 1 1 1 U 1 1 7 8 Wsh; BLM: S;<br />

USFWS:C<br />

Papaveraceae Arctomecon humilis Dwarf bearclaw<br />

poppy<br />

2 1 1 1 1 1 1 8 8 Wsh; USFWS:E<br />

Polemoniaceae<br />

Gilia caespitosa<br />

(Aliciella c.)<br />

Rabbit Valley<br />

gilia<br />

2 1 1 1 0 1 1 7 7 Way; BLM: S;<br />

USFS: S<br />

Ranunculaceae<br />

Ranunculus aestivalis<br />

(R. acris var. aestivalis)<br />

Autumn buttercup 2 1 1 1 1 1 1 8 8 Grf; USFWS:E<br />

Gibbens’ beardtongue<br />

Scrophulariaceae<br />

Penstemon gibbensii<br />

2 1 1 1 U 1 U 6 8 Dag; BLM: S<br />

Penstemon grahamii<br />

Graham’s penstemon<br />

2 1 1 1 1 1 1 8 8 Crb, Uin; BLM:<br />

S; USFWS:P<br />

Penstemon scariosus<br />

var. albifluvis<br />

White River penstemon<br />

2 1 1 1 0 1 1 7 7 Uin; BLM: S;<br />

USFWS: C<br />

Violaceae Viola clauseniana Clausen’s violet 2 1 1 1 1 U 1 7 8 Wsh<br />

209


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species<br />

The following table lists 119 species scored as High Priority for conservation attention in <strong>Utah</strong> based on the Wyoming<br />

protocol ranking system. Species are listed alphabetically by family and scientific name, with synonyms in parentheses.<br />

See text for an explanation of the seven ranking criteria and scoring methods used to derive the minimum<br />

and potential scores. County codes are explained in Table 3. Legal Status: Bureau of Land Management (BLM) and<br />

US Forest Service (USFS) Sensitive = S; US Fish and Wildlife Service (USFWS) Candidate = C, Endangered = E,<br />

Proposed = P; Threatened = T.<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Apiaceae<br />

(Umbelliferae)<br />

Cymopterus coulteri<br />

Two-leaf springparsley<br />

2 1 U 1 0 1 U 5 7 Jub, Snp, Sev,<br />

Toe<br />

Cymopterus higginsii<br />

Higgins’ springparsley<br />

2 1 1 1 0 1 0 6 6 Kan<br />

Lomatium latilobum<br />

Canyonlands<br />

lomatium<br />

2 1 1 1 0 1 U 6 7 Grn, Snj; BLM:<br />

S; USFS: S<br />

Lomatium scabrum var.<br />

tripinnatum<br />

Virgin lomatium 2 1 U 1 0 1 U 5 7 Wsh<br />

Apocynaceae<br />

Cycladenia humilis var.<br />

jonesii<br />

(C. jonesii)<br />

Jones’ cycladenia 1 1 1 1 1 1 0 6 6 Emr, Grf, Grn,<br />

Kan; USFWS:T<br />

Asclepiadaceae Asclepias welshii Welsh’s milkweed 2 1 1 1 0 1 0 6 6 Kan; USFWS:T<br />

Asteraceae<br />

(Compositae)<br />

Ambrosia x sandersonii<br />

(Hymenoclea s.)<br />

Chrysothamnus nauseosus<br />

var. glareosus<br />

(Ericameria nauseosa<br />

var. glareosa)<br />

Cirsium virginense<br />

(included in C. mohavense<br />

in FNA)<br />

Enceliopsis nudicaulis<br />

var. bairdii<br />

Sanderson’s bursage<br />

Marysvale rabbitbrush<br />

2 1 1 0 1 1 U 6 7 Wsh<br />

2 1 1 1 0 U 1 6 7 Piu<br />

Virgin thistle 1 1 1 1 0 1 1 6 6 Wsh; BLM: S<br />

Baird’s nakedstem 2 1 1 1 0 1 1 7 7 Wsh<br />

Erigeron higginsii<br />

(included in E. canaani<br />

in FNA)<br />

Higgins’ daisy 2 1 1 1 0 1 U 6 7 Wsh<br />

Erigeron kachinensis Kachina daisy 2 1 1 1 0 1 U 6 7 Snj; BLM: S;<br />

USFS: S; originally<br />

on Watch<br />

list<br />

Erigeon vagus var.<br />

madsenii<br />

Madsen’s daisy 2 1 1 1 1 0 U 6 7 Grf, Irn, Kan<br />

Haplopappus armerioides<br />

var. gramineus<br />

(Stenotus a. var. g.)<br />

Grass goldenweed 2 1 U 1 0 1 U 5 7 Dch, Uit<br />

210


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Asteraceae<br />

(Compositae)<br />

Haplopappus lignumviridis<br />

(Ericameria l.)<br />

Haplopappus scopulorum<br />

var. canonis<br />

Senecio castoreus<br />

(Packera c.)<br />

Senecio malmstenii<br />

(Packera m.)<br />

Senecio musiniensis<br />

(Packera m.)<br />

Thelesperma subnudum<br />

var. maliterrimum<br />

(T. pubescens)<br />

Townsendia goodrichii<br />

Townsendia jonesii var.<br />

lutea (included in T.<br />

aprica by some authors)<br />

Townsendia strigosa var.<br />

prolixa<br />

Viguiera soliceps<br />

(Heliomeris soliceps)<br />

Xylorhiza cronquistii<br />

Greenwood<br />

goldenbush<br />

Canyon spindly<br />

goldenbush<br />

Beaver Mountain<br />

groundsel<br />

2 1 1 1 1 0 U 6 7 Sev; BLM: S<br />

2 1 1 1 0 U U 5 7 Snj<br />

2 1 1 1 0 1 U 6 7 Bvr, Piu; USFS:<br />

S<br />

Podunk groundsel 2 1 1 1 1 0 U 6 7 Grf, Irn, Kan;<br />

USFS: S<br />

Musinea groundsel 2 1 1 1 U 0 U 5 7 Snp; USFS: S<br />

Uinta greenthread 2 1 1 1 0 U U 5 7 Dch, Uin; BLM:<br />

S; USFS: S<br />

Goodrich’s townsendia<br />

2 1 1 1 0 U U 5 7 Dch, Uin<br />

Sigurd townsendia 2 1 1 1 0 1 U 6 7 Jub, Piu, Sev;<br />

BLM: S; USFS:<br />

S<br />

Strigose townsendia<br />

2 1 U 1 0 1 U 5 7 Dch, Grn; BLM:<br />

S<br />

Tropic goldeneye 2 1 0 1 1 1 0 6 6 Kan<br />

Cronquist’s woodyaster<br />

2 1 1 1 1 0 U 6 7 Grf, Kan<br />

Xylorhiza glabriuscula<br />

var. linearifolia<br />

Boraginaceae Cryptantha grahamii Graham’s<br />

cryptanth<br />

Cryptantha semiglabra<br />

Moab woodyaster 2 1 U 1 0 1 U 5 7 Grf, Grn, Snj,<br />

Way<br />

Pipe Spring<br />

cryptanth<br />

2 1 U 1 0 1 U 5 7 Dch, Uin; BLM:<br />

S<br />

2 1 1 1 0 1 U 6 7 Wsh?<br />

Brassicaceae<br />

(Cruciferae)<br />

Arabis falcatoria<br />

(Boechera falcatoria)<br />

Arabis harrisonii<br />

(Boechera harrisonii)<br />

Draba ramulosa<br />

Falcate rockcress 2 1 1 1 0 U U 5 7 Box, Jub; USFS:<br />

S<br />

Harrison’s rockcress<br />

Belknap Peak<br />

draba<br />

2 1 1 1 0 U U 5 7 Uta<br />

2 1 1 1 0 1 U 6 7 Bvr, Piu; USFS:<br />

S<br />

Draba sobolifera Creeping draba 2 1 1 1 0 1 U 6 7 Bvr, Piu; USFS:<br />

S<br />

Lepidium integrifolium<br />

Entire-leaf pepperwort<br />

1 1 1 1 0 1 1 6 6 Bvr, Rch, Snp,<br />

Sev, Uin<br />

211


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Brassicaceae<br />

(Cruciferae)<br />

Cactaceae<br />

Capparaceae<br />

(Cleomaceae)<br />

Lepidium montanum var.<br />

alpinum<br />

Lepidium montanum var.<br />

stellae<br />

Lepidium ostleri<br />

Physaria chambersii var.<br />

canaanii<br />

Physaria grahamii<br />

(some authors include P.<br />

acutifolia vars. repanda<br />

& purpurea)<br />

Physaria rubicundula<br />

var. tumulosa<br />

(P. t., Lesquerella t.)<br />

Ferocactus acanthodes<br />

(F. cylindraceus)<br />

Pediocactus sileri<br />

Sclerocactus<br />

wetlandicus<br />

(S. whipplei var.<br />

glaucus)<br />

Cleomella hillmanii var.<br />

goodrichii<br />

(C. palmeriana var. g.)<br />

Chenopodiaceae Krascheninnikovia<br />

lanata var. ruinina<br />

Crassulaceae<br />

Dudleya pulverulenta<br />

var. arizonica<br />

(D. arizonica)<br />

Wasatch pepperwort<br />

Stella’s pepperwort<br />

Ostler’s pepperwort<br />

Canaan Peak twinpod<br />

2 1 1 1 0 U U 5 7 Slt; USFS: S<br />

1 1 1 1 1 1 U 6 7 Grf, Kan<br />

2 1 1 1 0 1 U 6 7 Bvr; BLM: S;<br />

USFWS: C<br />

2 1 1 1 0 U U 5 7 Grf<br />

Graham’s twinpod 2 1 1 1 0 U U 5 7 Dch, Grn, Uin,<br />

Uta, Was<br />

Kodachrome bladderpod<br />

Desert barrel cactus<br />

Siler’s pincushion<br />

cactus<br />

Uinta Basin hookless<br />

cactus<br />

Goodrich’s stinkweed<br />

Ruin Park winterfat<br />

Arizona liveforever<br />

2 1 1 1 0 1 U 6 7 Kan; USFWS:E<br />

1 1 1 1 0 1 1 6 6 Wsh; originally<br />

on Watch list<br />

1 1 1 1 0 1 1 6 6 Kan, Wsh;<br />

USFWS:T<br />

2 0 1 1 0 1 1 6 6 Dch, Uin;<br />

USFWS:T<br />

(formerly on<br />

ExH list)<br />

2 1 1 1 0 1 U 6 7 Uin; BLM: S<br />

2 1 1 1 0 1 U 6 7 Grn, Snj<br />

1 1 1 1 0 1 1 6 6 Wsh<br />

Cuscutaceae Cuscuta warneri Warner’s dodder 1 1 1 U 1 1 1 6 7 Mil; may be extirpated<br />

in UT<br />

Cyperaceae<br />

Fabaceae<br />

(Leguminosae)<br />

Carex haysii<br />

(included in C. curatorum<br />

by some authors)<br />

Hays’ sedge 2 1 1 1 0 1 U 6 7 Wsh; originally<br />

on Watch list<br />

Astragalus ampullarius Gumbo milkvetch 1 1 1 1 1 1 U 6 7 Kan, Wsh; BLM:<br />

S<br />

Astragalus cronquistii<br />

Cronquist’s milkvetch<br />

2 1 1 1 0 1 0 6 6 Snj; BLM: S<br />

Astragalus cutleri Cutler’s milkvetch 2 1 1 1 0 U U 5 7 Snj<br />

Astragalus desereticus Deseret milkvetch 2 1 1 1 U 1 0 6 7 Uta; USFWS:T<br />

212


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Fabaceae<br />

(Leguminosae)<br />

Astragalus diversifolius<br />

Meadow milkvetch<br />

1 1 1 1 0 1 1 6 6 Jub, Toe; USFS:<br />

S<br />

Astragalus equisolensis<br />

(A. desperatus var.<br />

neeseae)<br />

Horseshoe milkvetch<br />

2 1 1 1 0 1 U 6 7 Uin; BLM: S<br />

Astragalus hamiltonii<br />

Hamilton’s milkvetch<br />

2 1 1 1 0 1 U 6 7 Uin; BLM: S<br />

Astragalus harrisonii<br />

Harrison’s milkvetch<br />

2 1 1 1 0 1 U 6 7 Grf, Way<br />

Astragalus loanus<br />

Glenwood milkvetch<br />

2 1 1 1 0 U U 5 7 Sev; BLM: S<br />

Astragalus sabulosus<br />

var. sabulosus<br />

Astragalus sabulosus<br />

var. vehiculus<br />

Cisco milkvetch 2 1 1 1 0 1 U 6 7 Grn; BLM: S<br />

Stage milkvetch 2 1 1 1 0 1 U 6 7 Grn; BLM: S<br />

Astragalus serpens Plateau milkvetch 2 1 1 1 0 U U 5 7 Piu, Sev, Way<br />

Astragalus striatiflorus<br />

Escarpment milkvetch<br />

2 1 1 1 0 1 U 6 7 Kan, Wsh; BLM:<br />

S<br />

Astragalus welshii Welsh’s milkvetch 2 1 1 1 0 U U 5 7 Grf, Irn, Kan,<br />

Mill Piu, Way;<br />

BLM: S<br />

Trifolium friscanum<br />

(T. andersonii var. f.)<br />

Frisco clover 2 1 1 1 0 1 U 6 7 Bvr; BLM: S;<br />

USFWS: C<br />

Fagaceae<br />

Quercus gambelii var.<br />

bonina<br />

Goodhope oak 2 1 1 0 1 1 U 6 7 Snj<br />

Fumariaceae<br />

Corydalis caseana var.<br />

brachycarpa<br />

Gentianaceae Frasera ackermaniae Ackerman’s frasera<br />

Hydrangeaceae<br />

(Saxifragaceae)<br />

Jamesia americana var.<br />

macrocalyx<br />

Hydrophyllaceae Phacelia argylensis<br />

Case’s corydalis 2 1 1 0 0 1 1 6 6 Slt, Uta, Was,<br />

Web; USFS: S<br />

2 1 1 1 U 0 U 5 7 Uin; BLM: S<br />

Wasatch jamesia 2 1 1 1 0 U U 5 7 Jub, Slt, Uta,<br />

Was; USFS: S<br />

Argyle Canyon<br />

phacelia<br />

2 1 1 1 0 1 U 6 7 Dch; BLM: S<br />

Phacelia cephalotes Chinle phacelia 1 1 1 1 1 1 U 6 7 Kan, Snj, Wsh<br />

Phacelia cronquistiana<br />

Cronquist’s phacelia<br />

1 1 1 1 1 1 U 6 7 Kan; BLM: S<br />

Phacelia demissa var.<br />

heterotricha<br />

Brittle phacelia 2 1 0 1 1 1 U 6 7 Piu, Sev, Way<br />

213


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Hydrophyllaceae<br />

Phacelia demissa var.<br />

minor<br />

Brittle phacelia 2 1 0 1 1 1 U 6 7 Dch, Uin<br />

Phacelia indecora Bluff phacelia 2 1 1 1 0 1 U 6 7 Snj; BLM: S;<br />

originally on<br />

Watch list<br />

Phacelia pulchella var.<br />

atwoodii<br />

Phacelia pulchella var.<br />

gooddingii<br />

Phacelia sabulonum<br />

(P. pulchella var.<br />

sabulonum)<br />

Atwood’s pretty<br />

phacelia<br />

Goodding’s pretty<br />

phacelia<br />

2 1 0 1 1 1 0 6 6 Kan; BLM: S<br />

1 1 1 1 1 1 U 6 7 Wsh<br />

Tompkins phacelia 2 1 0 1 1 1 U 6 7 Grf, Kan<br />

Loasaceae Mentzelia shultziorum Shultz’s stickleaf 2 1 1 1 0 1 U 6 7 Grn; BLM: S<br />

Malvaceae<br />

Petalonyx parryi<br />

Sphaeralcea fumariensis<br />

(S. grossulariifolia var.<br />

fumariensis)<br />

Sphaeralcea janeae<br />

Parry’s sandpaperplant<br />

Smoky Mountain<br />

globemallow<br />

Jane’s globemallow<br />

Sphaeralcea psoraloides Scurfpea globemallow<br />

Onagraceae Camissonia exilis Meager camissonia<br />

Oenothera caespitosa<br />

var. stellae<br />

Stella’s eveningprimrose<br />

1 1 1 1 0 1 1 6 6 Wsh; BLM: S<br />

2 1 1 1 0 1 U 6 7 Kan: BLM: S<br />

2 1 1 1 0 U U 5 7 Grn, Snj, Way;<br />

BLM: S<br />

2 1 1 1 0 U U 5 7 Emr, Grn, Way;<br />

BLM: S<br />

1 1 1 1 1 1 U 6 7 Kan<br />

2 1 1 1 0 1 U 6 7 Emr, Grf, Kan,<br />

Snp<br />

Oenothera murdockii<br />

Murdock’s evening-primrose<br />

2 1 1 1 0 1 U 6 7 Kan, Wsh; BLM:<br />

S<br />

Ophioglossaceae Botrychium lineare Slender moonwort 1 1 1 U 1 1 U 5 7 Slt; USFS: S<br />

Orchidaceae<br />

Cypripedium calceolus<br />

var. parviflorum<br />

Spiranthes diluvialis<br />

(S. romanzoffiana var.<br />

d.)<br />

Large yellow ladies-slipper<br />

1 1 1 0 1 1 1 6 6 Cch, Grn, Slt,<br />

Sum, Uta, Web;<br />

USFS: S<br />

Ute ladies-tresses 1 1 0 1 1 1 1 6 6 Cch, Dag, Dch,<br />

Grf, Slt, Toe,<br />

Uin, Uta, Way,<br />

Web; USFWS:T<br />

Poaceae Elymus simplex Alkali wildrye 1 1 1 1 U 1 1 6 7 Dag<br />

Polemoniaceae<br />

Gilia imperialis<br />

(G. latifolia var. i.)<br />

Cataract gilia 2 1 1 1 0 U U 5 7 Emr, Grf, Kan,<br />

Snj, Way<br />

214


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Polemoniaceae<br />

Polygonaceae<br />

Gilia tenuis<br />

(Aliciella t.)<br />

Ipomopsis congesta var.<br />

ochroleuca<br />

Phlox hoodii var.<br />

madsenii<br />

Eriogonum brevicaule<br />

var. huberi<br />

(included in var. laxifolium<br />

by Holmgren et<br />

al. <strong>2012</strong>)<br />

Eriogonum brevicaule<br />

var. mitophyllum<br />

(E. mitophyllum)<br />

Eriogonum brevicaule<br />

var. promiscuum<br />

(included in var. laxifolium<br />

by Holmgren et<br />

al. <strong>2012</strong>)<br />

Eriogonum corymbosum<br />

var. cronquistii<br />

(E. cronquistii)<br />

Eriogonum corymbosum<br />

var. heilii<br />

Eriogonum corymbosum<br />

var. matthewsiae<br />

(included in var. albiflorum<br />

by Holmgren et al.<br />

<strong>2012</strong>)<br />

Eriogonum corymbosum<br />

var. smithii<br />

(E. smithii)<br />

Mussentuchit gilia 2 1 1 1 0 U U 5 7 Emr, Sev; BLM:<br />

S<br />

Arapien gilia 2 1 U 1 0 1 U 5 7 Snp, Sev<br />

Madsen’s carpet<br />

phlox<br />

Huber’s wild<br />

buckwheat<br />

Lost Creek wild<br />

buckwheat<br />

Mount Bartles<br />

wild buckwheat<br />

Cronqist’s wild<br />

buckwheat<br />

Heil’s wild buckwheat<br />

Springdale wild<br />

buckwheat<br />

Flat top wild buckwheat<br />

2 1 1 1 0 U U 5 7 Way<br />

2 1 1 1 0 U U 5 7 Dch<br />

2 1 1 1 0 1 U 6 7 Sev; BLM: S<br />

2 1 1 1 0 U U 5 7 Crb<br />

2 1 1 1 0 U U 5 7 Grf<br />

2 1 1 1 0 1 U 6 7 Way<br />

2 1 1 1 0 1 U 6 7 Wsh<br />

2 1 1 1 0 U U 5 7 Emr, Way;<br />

BLM: S<br />

Eriogonum esmeraldense<br />

var. tayei<br />

(Included in var. esmeraldense<br />

by Holmgren et<br />

al. <strong>2012</strong>)<br />

Eriogonum nummulare<br />

var. ammophilum<br />

(E. ammophilum)<br />

Eriogonum racemosum<br />

var. nobilis<br />

(included in E. zionis by<br />

Holmgren et al. <strong>2012</strong>)<br />

Taye’s wild buckwheat<br />

Ibex wild buckwheat<br />

Bluff wild buckwheat<br />

2 1 1 1 0 U U 5 7 Sev<br />

2 1 1 1 0 U U 5 7 Mil; BLM: S<br />

2 1 1 1 0 U U 5 7 Kan, Snj; BLM:<br />

S<br />

215


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 2. UNPS Rare <strong>Plant</strong> List: High Priority Species, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Polygonaceae Eriogonum soredium Frisco wild buckwheat<br />

Portulacaceae<br />

(Montiaceae)<br />

Primulaceae<br />

Talinum thompsonii<br />

(Phemeranthus validulus)<br />

Dodecatheon dentatum<br />

var. utahense<br />

(D. utahense)<br />

Primula domensis<br />

Thompson’s<br />

talinum<br />

Hooker’s shooting-star<br />

House Range<br />

primrose<br />

2 1 1 1 0 1 U 6 7 Bvr; BLM: S<br />

2 1 1 1 0 U U 5 7 Emr; BLM: S<br />

2 1 1 1 1 0 0 6 6 Slt; USFS: S<br />

2 1 1 1 0 U U 5 7 Mil; BLM: S<br />

Primula maguirei<br />

Maguire’s primrose<br />

2 1 1 1 1 0 U 6 7 Cch; USFWS:T<br />

Ranunculaceae<br />

Aquilegia holmgrenii<br />

(formerly included in A.<br />

elegantula)<br />

Holmgren’s columbine<br />

2 1 1 1 0 U U 5 7 Grf<br />

Aquilegia rubicunda<br />

Link Trail columbine<br />

2 1 1 1 0 U U 5 7 Emr, Sev; USFS:<br />

S<br />

Aquilegia scopulorum<br />

var. goodrichii<br />

Goodrich’s columbine<br />

2 1 1 1 0 U U 5 7 Dch; BLM: S<br />

Rosaceae<br />

Ivesia shockleyi var.<br />

ostleri<br />

Shockley’s ivesia 2 1 1 1 0 U U 5 7 Bvr; BLM: S<br />

Aquarius paintbrush<br />

Scrophulariaceae<br />

Castilleja aquariensis<br />

Castilleja parvula var.<br />

revealii<br />

Penstemon flowersii<br />

Penstemon goodrichii<br />

Penstemon x jonesii<br />

Reveal’s paintbrush<br />

Flowers’ penstemon<br />

Goodrich’s penstemon<br />

Fuchsia penstemon<br />

2 1 0 1 1 1 0 6 6 Grf; USFS: S<br />

2 1 1 1 1 0 U 6 7 Grf, Irn, Kan;<br />

USFS: S<br />

2 1 1 1 0 1 U 6 7 Dch, Uin<br />

2 1 1 1 0 1 U 6 7 Dch, Uin; BLM:<br />

S<br />

2 1 1 0 1 U U 5 7 Kan, Wsh<br />

Penstemon pinorum Pinyon penstemon 2 1 1 1 0 1 U 6 7 Irn; BLM: S;<br />

USFS: S<br />

Penstemon tidestromii<br />

(includes P. leptanthus)<br />

Tidestrom’s penstemon<br />

2 1 1 1 0 1 U 6 7 Jub, Snp, Uta<br />

Penstemon wardii Ward’s penstemon 2 1 1 1 0 1 U 6 7 Mil, Piu, Snp,<br />

Sev; BLM: S;<br />

USFS: S<br />

Violaceae Viola beckwithii Beckwith’s violet 1 1 1 U U 1 1 5 7 Box, Cch, Slt,<br />

Uta<br />

216


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List<br />

The following table lists 264 species on the Watch List for potential conservation attention in <strong>Utah</strong> based on the<br />

Wyoming protocol ranking system. Species are listed alphabetically by family and scientific name, with synonyms in<br />

parentheses. See text for an explanation of the seven ranking criteria and scoring methods used to derive the minimum<br />

and potential scores. County codes are explained in Table 3. Legal Status: Bureau of Land Management<br />

(BLM) and US Forest Service (USFS) Sensitive = S; US Fish and Wildlife Service (USFWS) Candidate = C, Endangered<br />

= E, Proposed = P; Threatened = T.<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Adoxaceae Adoxa moschatellina Moschatel 1 1 1 1 0 U 1 5 6 Snj<br />

Agavaceae<br />

Apiaceae<br />

(Umbelliferae)<br />

Agave utahensis var.<br />

utahensis<br />

<strong>Utah</strong> century plant 1 1 1 1 0 1 U 5 6 Wsh<br />

Nolina microcarpa Beargrass 1 1 1 0 0 1 1 5 5 Wsh<br />

Yucca kanabensis<br />

(Y. angustissima var. k.)<br />

Kanab yucca 1 1 1 1 1 0 U 5 6 Kan, Wsh<br />

Yucca schidigera Splinter yucca 1 1 1 0 0 1 1 5 5 Wsh; originally<br />

High priority<br />

Yucca toftiae<br />

(Y. angustissima var.<br />

toftiae)<br />

Toft’s yucca 2 1 1 1 0 0 U 5 6 Grf, Kan, Snj;<br />

originally High<br />

priority<br />

Angelica wheeleri <strong>Utah</strong> angelica 1 1 1 0 0 1 1 5 5 Cch, Jub, Piu,<br />

Slt, Sev, Uta;<br />

USFS:S<br />

Cymopterus acaulis var.<br />

parvus<br />

Cymopterus beckii<br />

Cymopterus evertii<br />

Cymopterus minimus<br />

Cymopterus trotteri<br />

(Oreoxis trotteri)<br />

Lomatium graveolens<br />

var. clarkii<br />

Small springparsley<br />

Beck’s springparsley<br />

Evert’s springparsley<br />

Least springparsley<br />

Trotter’s springparsley<br />

2 1 U 1 0 0 U 4 6 Mil, Toe<br />

1 1 0 1 0 1 1 5 5 Kan, Snj, Way;<br />

BLM: S; USFS:<br />

S<br />

1 1 1 1 0 U U 4 6 Uin<br />

2 1 1 1 0 0 U 5 6 Grf, Irn, Kan;<br />

USFS: S<br />

2 1 1 1 0 0 U 5 6 Grn; BLM: S<br />

Clark’s lomatium 2 1 1 1 0 0 0 5 5 Wsh<br />

Lomatium junceum Rush lomatium 2 1 1 0 0 U U 4 6 Emr, Grf, Sev,<br />

Way<br />

Musineon lineare <strong>Utah</strong> musineon 2 1 1 1 0 0 U 5 6 Box, Cch<br />

Asclepiadaceae<br />

Asclepias cutleri Cutler’s milkweed 1 1 1 1 0 1 U 5 6 Grn, Snj<br />

Cynanchum utahense Swallow-wort 1 1 1 0 0 1 1 5 5 Wsh<br />

Asteraceae<br />

(Compositae)<br />

Artemisia campestris<br />

var. petiolata<br />

Petiolate wormwood<br />

2 1 1 0 0 U U 4 6 Dch; USFS: S<br />

217


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Asteraceae<br />

(Compositae)<br />

Artemisia nova var.<br />

duchesnicola<br />

Aster kingii var.<br />

barnebyana<br />

(Tonestus k. var. b., Herrickia<br />

k. var. b.)<br />

Aster kingii var. kingii<br />

(Tonestus k. var. k., Herrickia<br />

k. var. k.)<br />

Aster welshii<br />

(Symphyotrichum w.)<br />

Baccharis viminea var.<br />

atwoodii<br />

Chrysopsis jonesii<br />

(Heterotheca jonesii)<br />

Chrysothamnus nauseosus<br />

var. iridis<br />

(Ericameria nauseosa<br />

var. i.)<br />

Chrysothamnus nauseosus<br />

var. psilocarpus<br />

(Ericameria nauseosa<br />

var. psilocarpa)<br />

Cirsium eatonii var.<br />

harrisonii<br />

Duchesne sagebrush<br />

2 1 1 1 0 0 U 5 6 Uin<br />

Barneby’s aster 2 1 1 1 0 0 U 5 6 Jub, Mil; USFS:<br />

S<br />

King’s aster 2 1 1 1 0 0 U 5 6 Slt, Uta; USFS:<br />

S<br />

Welsh’s aster 1 1 1 1 0 1 U 5 6 Bvr, Dch, Grf,<br />

Irn, Kan, Piu,<br />

Sum, Uta, Wsh,<br />

Way<br />

Atwood’s seepwillow<br />

Jones’ goldenaster<br />

Rainbow rabbitbrush<br />

Huntington rabbitbrush<br />

2 1 U 0 0 1 U 4 6 Emr, Grn, Snj<br />

2 1 1 1 0 0 0 5 5 Grf, Kan, Wsh;<br />

USFS: S<br />

2 1 0 1 0 1 U 5 6 Snp, Sev<br />

2 1 1 0 0 U U 4 6 Crb, Dch, Emr,<br />

Sev, Was<br />

Harrison’s thistle 2 1 U 1 0 0 U 4 6 Bvr, Piu<br />

Cirsium joannae Joanna’s thistle 2 1 1 1 0 0 U 5 6 Kan, Wsh<br />

Cirsium murdockii<br />

(C. eatonii var. m.)<br />

Murdock’s thistle 2 1 0 1 0 U U 4 6 Dag, Dch, Uin<br />

Cirsium ownbeyi Ownbey’s thistle 1 1 1 1 0 1 U 5 6 Dag, Uin<br />

Enceliopsis argophylla<br />

Silverleaf enceliopsis<br />

1 U U 1 0 1 1 4 6 Wsh<br />

Erigeron arenarioides Wasatch daisy 2 1 1 1 0 0 0 5 5 Box, Slt, Toe,<br />

Uta, Web<br />

Erigeron canaani Canaan daisy 2 1 1 1 0 0 0 5 5 Kan, Wsh<br />

Erigeron carringtoniae<br />

(included in E. untermannii<br />

in FNA)<br />

Carrington’s daisy 2 1 1 1 0 0 U 5 6 Emr, Snp; USFS:<br />

S<br />

Erigeron cronquistii Cronquist’s daisy 2 1 1 1 0 0 U 5 6 Cch; BLM: S;<br />

USFS: S<br />

218


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Asteraceae<br />

(Compositae)<br />

Erigeron garrettii Garrett’s daisy 2 1 1 1 0 0 U 5 6 Slt, Uta, Was;<br />

USFS:S<br />

Erigeron goodrichii Goodrich’s daisy 2 1 U 1 0 0 U 4 6 Dag, Dch, Sum?,<br />

Uin, Uta<br />

Erigeron huberi<br />

(included in E. radicatus<br />

in FNA)<br />

Erigeron maguirei<br />

(includes var. harrisonii)<br />

Huber’s daisy 2 1 U 1 0 0 U 4 6 Dch<br />

Maguire’s daisy 2 1 1 1 0 0 0 5 5 Emr, Way;<br />

BLM: S;<br />

USFWS former<br />

T<br />

Erigeron religiosus Religious daisy 2 0 1 1 0 1 0 5 5 Grf, Kan, Snj,<br />

Wsh<br />

Erigeron sionis<br />

(includes vars. sionis &<br />

trilobatus)<br />

Zion daisy 2 1 1 1 0 0 0 5 5 Grf, Irn, Kan,<br />

Wsh<br />

Erigeron untermannii Untermann’s daisy 2 1 0 1 0 U U 4 6 Dch; BLM: S;<br />

USFS: S<br />

Erigeron ursinus var.<br />

meyerae<br />

Meyer’s daisy 2 1 1 0 U 0 U 4 6 Wsh<br />

Erigeron zothecinus Alcove daisy 2 1 1 1 0 0 U 5 6 Grf, Grn, Kan,<br />

Snj<br />

Geraea canescens Desert sunflower 1 1 1 0 0 1 1 5 5 Wsh<br />

Gutierrezia pomariensis Orchard snakeweed<br />

Haplopappus racemosus<br />

var. sessiliflorus<br />

(Pyrrocoma racemosa<br />

var. sessiliflora)<br />

Haplopappus zionis<br />

(Ericameria z.)<br />

Hymenoxys helenioides<br />

(Picradenia helenioides)<br />

Racemose goldenweed<br />

Cedar Breaks<br />

goldenweed<br />

Sneezeweed hymenoxys<br />

2 1 U 1 0 0 U 4 6 Dch, Uin<br />

1 1 1 1 0 U U 4 6 Mil<br />

2 1 1 1 0 0 U 5 6 Grf, Irn, Kan;<br />

BLM: S<br />

1 1 1 0 1 U U 4 6 Crb, Emr, Grf,<br />

Snp, Sev, Way<br />

Hymenoxys lapidicola Rock hymenoxys 2 1 1 1 0 0 0 5 5 Uin; BLM: S<br />

Hymenoxys lemmonii Alkali hymenoxys 1 1 1 1 0 U U 4 6 Uin<br />

Layia platyglossa var.<br />

breviseta<br />

Lepidospartum latisquamum<br />

Perityle emoryi<br />

Coastal tidytips 1 1 1 1 0 U 1 5 6 Snj<br />

Nevada broom 1 1 U 1 0 1 U 4 6 Mil<br />

Emory’s rockdaisy<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

219


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Asteraceae<br />

(Compositae)<br />

Perityle specuicola Alcove rock-daisy 2 1 1 1 0 0 U 5 6 Grn, Snj; BLM:<br />

S<br />

Peucephyllum schottii Pygmy-cedar 1 1 1 1 0 1 U 5 6 Wsh<br />

Platyschkuhria integrifolia<br />

var. oblongifolia<br />

Senecio dimorphophyllus<br />

var. intermedius<br />

(Packera dimorphophylla<br />

var. intermedia)<br />

Senecio fremontii var.<br />

inexpectans<br />

San Juan bahia 1 1 1 1 0 U U 4 6 Snj<br />

La Sal groundsel 2 1 1 0 0 U U 4 6 Dch, Grn, Snj,<br />

Snp, Sum<br />

Unexpected<br />

groundsel<br />

2 1 1 1 0 0 U 5 6 Grn, Snj<br />

Senecio werneriifolius<br />

var. barkleyi<br />

Barkley’s groundsel<br />

220<br />

2 1 1 1 0 0 U 5 6 Grf, Kan<br />

Solidago spectabilis Nevada goldenrod 1 1 U 1 0 1 U 4 6 Mil, Wsh<br />

Sphaeromeria ruthiae<br />

(Artmeisia ruthiae)<br />

Stephanomeria tenuifolia<br />

var. myrioclada<br />

(S. minor var. myrioclada)<br />

Stephanomeria tenuifolia<br />

var. uintahensis<br />

(S. minor var. u.)<br />

Townsendia beamanii<br />

Ruth’s chickensage<br />

Slender wirelettuce<br />

2 1 1 1 0 0 0 5 5 Kan, Wsh<br />

1 1 1 1 0 U U 4 6 Box<br />

Uinta wire-lettuce 2 1 1 0 0 U U 4 6 Uin<br />

Beaman’s townsendia<br />

Townsendia condensata Cushion townsendia<br />

Townsendia mensana<br />

Townsendia montana<br />

var. caelilinesis<br />

Townsendia montana<br />

var. minima<br />

Xylorhiza confertifolia<br />

Plateau townsendia<br />

Skyline townsendia<br />

Bryce Canyon<br />

townsendia<br />

Henrieville<br />

woody-aster<br />

Boraginaceae Cryptantha barnebyi Barneby’s cryptanth<br />

2 1 1 0 U 0 U 4 6 Snj; BLM: S<br />

1 1 1 1 0 U U 4 6 Bvr, Piu<br />

2 0 U 1 0 1 U 4 6 Dch, Uin<br />

2 1 0 1 0 U U 4 6 Dch, Snp, Was<br />

2 1 1 1 0 0 U 5 6 Grf, Irn, Kan,<br />

Wsh<br />

2 1 0 1 0 1 U 5 6 Grf, Kan, Way<br />

2 1 0 1 0 1 U 5 6 Uin; BLM: S<br />

Cryptantha compacta Mound cryptanth 2 1 0 1 0 U U 4 6 Bvr, Mil, Toe;<br />

BLM: S<br />

Cryptantha creutzfeldtii<br />

Creutzfeldt’s<br />

cryptanth<br />

2 1 0 1 0 1 U 5 6 Crb, Emr; BLM:<br />

S; USFS: S


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Boraginaceae Cryptantha elata Tall cryptanth 1 1 1 1 0 U U 4 6 Grn<br />

Brassicaceae<br />

(Cruciferae)<br />

Cryptantha johnstonii<br />

Cryptantha jonesiana<br />

Cryptantha ochroleuca<br />

(included in C. compacta<br />

by some authors)<br />

Hackelia ibapensis<br />

Arabis shockleyi<br />

(Boechera s.)<br />

Arabis vivariensis<br />

(Boechera. fernaldiana<br />

ssp. v.)<br />

Descurainia pinnata var.<br />

paysonii<br />

(D. incisa var. p.)<br />

Johnston’s cryptanth<br />

San Rafael cryptanth<br />

Yellowish cryptanth<br />

Deep Creek stickseed<br />

Shockley’s rockcress<br />

2 1 1 0 0 1 U 5 6 Emr<br />

2 1 0 1 0 U U 4 6 Emr<br />

2 1 1 1 0 0 0 5 5 Grf; USFS: S<br />

2 1 1 1 0 0 U 5 6 Jub<br />

1 1 1 1 0 U U 4 6 Bvr, Jub, Mil,<br />

Toe<br />

Park rockcress 2 1 1 1 0 0 U 5 6 Uin; BLM: S<br />

Payson’s tansymustard<br />

1 1 1 1 0 U U 4 6 Grn, Snj, Uin<br />

Draba kassii Kass’ draba 2 1 1 1 0 0 U 5 6 Toe<br />

Draba maguirei var.<br />

burkei<br />

(D. burkei)<br />

Draba maguirei var.<br />

maguirei<br />

Lepidium huberi<br />

Lepidium montanum var.<br />

claronense<br />

Lepidium montanum var.<br />

heterophyllum<br />

Lepidium montanum var.<br />

neeseae<br />

Burke’s draba 2 1 1 1 0 0 U 5 6 Box, Mor, Web;<br />

USFS: S<br />

Maguire’s draba 2 1 U 1 0 0 U 4 6 Box, Cch, Web;<br />

USFS: S<br />

Huber’s pepperwort<br />

1 1 1 1 0 U U 4 6 Uin; BLM: S<br />

Claron pepperwort 2 1 1 1 0 0 U 5 6 Grf, Kan, Piu<br />

Cedar Canyon<br />

pepperwort<br />

Neese’s pepperwort<br />

2 1 1 0 0 1 U 5 6 Irn, Mil, Piu, Sev<br />

2 1 1 1 0 0 U 5 6 Grf; USFS: S<br />

Lepidium nanum Low pepperwort 1 1 1 1 0 U U 4 6 Toe<br />

Physaria acutifiolia var.<br />

purpurea (included in<br />

P. grahamii in FNA)<br />

Physaria arizonica<br />

(Lesquerella arizonica)<br />

Physaria chambersii var.<br />

sobolifera<br />

Purple twinpod 2 1 1 1 0 0 U 5 6 Emr, Grn, Sev,<br />

Way<br />

Arizona bladderpod<br />

1 1 1 1 0 U U 4 6 Grf, Kan, Wsh<br />

Claron twinpod 2 1 1 1 0 0 U 5 6 Grf<br />

Physaria floribunda Mesa twinpod 1 1 1 1 0 U U 4 6 Grn<br />

221


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Brassicaceae<br />

(Cruciferae)<br />

Physaria garrettii<br />

(Lesquerella garrettii)<br />

Thelypodiopsis ambigua<br />

var. erecta<br />

Thelypodiopsis sagittata<br />

var. ovalifolia<br />

(Thelypodium sagittatum<br />

var. ovalifolium)<br />

Garrett’s twinpod 2 1 1 1 0 0 U 5 6 Dav, Slt, Uta,<br />

Was; USFS: S<br />

Kanab thelypody 1 1 1 1 0 1 U 5 6 Kan, Wsh?;<br />

BLM: S<br />

Palmer’s thelypody<br />

1 1 1 1 0 U U 4 6 Grf, Irn, Jub,<br />

Kan, Mil<br />

Thelypodium flexuosum Zigzag thelypody 1 1 1 1 0 U U 4 6 Bvr, Toe<br />

Buddlejaceae Buddleja utahensis <strong>Utah</strong> butterflybush 1 1 1 1 0 1 0 5 5 Wsh<br />

Cactaceae<br />

Echinocereus triglochidiatus<br />

var. mojavensis<br />

(E. mojavensis)<br />

Mohave claretcup 1 1 1 0 0 1 1 5 5 Bvr, Mil, Wsh<br />

Mamillaria tetrancistra Pincushion cactus 1 1 1 0 0 1 1 5 5 Wsh<br />

Neolloydia johnsonii<br />

(Echinomastus j.)<br />

Opuntia echinocarpa<br />

(Cylindropuntia echinocarpa)<br />

Opuntia pulchella<br />

(Grusonia p.)<br />

Sclerocactus blainei<br />

Johnson’s neolloydia<br />

222<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

Pale cholla 1 1 1 0 0 1 1 5 5 Bvr?, Wsh<br />

Sand cholla 1 1 U 1 0 1 U 4 6 Box, Jub, Mil,<br />

Toe, Wsh?<br />

Blaine’s fishhook<br />

cactus<br />

1 1 U 1 0 1 1 5 6 Irn<br />

Caryophyllaceae Silene nachlingerae Jan’s catchfly 1 1 1 1 0 U U 4 6 Bvr; USFS: S<br />

Chenopodiaceae Atriplex gardneri var.<br />

bonnevillensis<br />

(A. bonnevillensis)<br />

Cuscutaceae<br />

Atriplex obovata<br />

Atriplex pleiantha<br />

(Proatriplex p.)<br />

Atriplex wolfii var.<br />

tenuissima<br />

Bonneville saltbush<br />

New Mexico saltbush<br />

Four Corners<br />

orach<br />

1 1 U 1 0 1 U 4 6 Jub, Mil<br />

1 1 U 1 0 1 U 4 6 Snj<br />

1 1 1 1 0 U U 4 6 Snj<br />

Slender orach 1 1 1 1 0 U U 4 6 Crb, Dch, Emr,<br />

Grf, Piu, Snp,<br />

Sev, Uin<br />

Corispermum welshii Welsh’s bugseed 1 1 U 1 0 1 U 4 6 Grf, Kan, Mil,<br />

Snj?<br />

Cuscuta applanata Winged dodder 1 1 1 0 1 U U 4 6 Wsh<br />

Cuscuta cuspidata Toothed dodder 1 1 1 0 1 U U 4 6 Slt, Uta, Web<br />

Cyperaceae Carex crawei Crawe’s sedge 1 1 1 1 0 U U 4 6 Kan<br />

Carex curatorum<br />

(C. scirpoidea var. c.)<br />

Canyonlands<br />

sedge<br />

1 1 1 1 0 U U 4 6 Kan, Snj, Uin


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Cyperaceae Carex diandra Lesser panicled<br />

sedge<br />

1 1 1 1 0 U U 4 6 Dch, Grf?<br />

Carex lasiocarpa Slender sedge 1 1 1 1 0 U U 4 6 Dag, Sev?, Uin<br />

Carex leptalea Bristly-stalk sedge 1 1 1 1 0 U U 4 6 Dag, Dch, Uin<br />

Carex livida Pale sedge 1 1 1 1 0 U U 4 6 Dch, Uin<br />

Carex microglochin Subulate sedge 1 1 1 1 0 U U 4 6 Dag, Dch, Emr<br />

Cladium californicum Saw-grass 1 1 U 1 0 U 1 4 6 Kan, Snj<br />

Lipocarpha aristulata<br />

(L. drummondii, Hemicarpha<br />

micrantha)<br />

Scirpus nevadensis<br />

(Amphiscirpus n.)<br />

Slender-rush 1 1 1 1 0 1 U 5 6 Kan<br />

Nevada bulrush 1 1 1 1 0 U U 4 6 Jub, Rch<br />

Euphorbiaceae Euphorbia nephradenia <strong>Utah</strong> spurge 1 1 1 1 0 1 U 5 6 Emr, Grf, Kan,<br />

Way; BLM: S<br />

Fabaceae<br />

(Leguminosae)<br />

Astragalus calycosus<br />

var. monophyllidus<br />

One-leaf milkvetch<br />

1 1 1 1 0 U U 4 6 Sev<br />

Astragalus chloodes Grass milkvetch 2 1 0 1 0 U U 4 6 Uin<br />

Astragalus concordius<br />

(formerly included in A.<br />

piutensis)<br />

Hairy-pod milkvetch<br />

2 1 U 0 0 1 U 4 6 Irn, Wsh<br />

Astragalus detritalis Debris milkvetch 1 1 1 1 0 1 U 5 6 Dch, Uin<br />

Astragalus henrimontanensis<br />

(A. argophyllus<br />

var. stocksii)<br />

Dana’s milkvetch 2 1 1 0 0 U U 4 6 Grf; USFS: S<br />

Astragalus jejunus<br />

Astragalus lentiginosus<br />

var. mokiacensis<br />

(A. m.; some authors<br />

include var. ursinus)<br />

Astragalus limnocharis<br />

var. limnocharis<br />

Astragalus limnocharis<br />

var. tabulaeus<br />

(included in A. montii by<br />

some authors)<br />

Starveling milkvetch<br />

1 1 1 1 0 U U 4 6 Rch; USFS: S<br />

Mokiak milkvetch 1 1 1 1 0 1 0 5 5 Wsh<br />

Navajo Lake milkvetch<br />

Table Cliff milkvetch<br />

2 1 1 1 0 0 U 5 6 Irn, Kan; USFS:<br />

S<br />

2 1 1 1 0 0 U 5 6 Grf; USFS: S<br />

Astragalus lutosus Dragon milkvetch 1 1 1 1 0 1 U 5 6 Dch, Uin, Uta,<br />

Was<br />

Astragalus malacoides<br />

Kaiparowits milkvetch<br />

2 1 1 1 0 0 U 5 6 Grf, Kan<br />

223


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Fabaceae<br />

(Leguminosae)<br />

Astragalus montii<br />

Astragalus monumentalis<br />

Heliotrope milkvetch<br />

Monument milkvetch<br />

2 1 0 1 0 U U 4 6 Snp, Sev;<br />

USFWS: T<br />

2 1 0 1 0 U U 4 6 Grf, Snj<br />

Astragalus naturitensis Naturita milkvetch 1 1 1 1 0 U U 4 6 Snj<br />

Astragalus piscator Fisher milkvetch 2 1 1 1 0 0 U 5 6 Grn, Snj, Way<br />

Astragalus saurinus<br />

Dinosaur milkvetch<br />

2 1 0 1 0 U U 4 6 Uin<br />

Astragalus uncialis Currant milkvetch 2 1 1 0 0 1 U 5 6 Mil; USFS: S<br />

Astragalus wetherillii<br />

Astragalus zionis var.<br />

vigulus (“A. tephrodes”)<br />

Hedysarum boreale var.<br />

gremiale<br />

Hedysarum occidentale<br />

var. canone<br />

Oxytropis besseyi var.<br />

obnapiformis<br />

Oxytropis oreophila var.<br />

jonesii<br />

Pediomelum aromaticum<br />

var. aromaticum<br />

Pediomelum aromaticum<br />

var. barnebyi<br />

Pediomelum aromaticum<br />

var. tuhyi<br />

Wetherill’s milkvetch<br />

224<br />

1 1 1 0 U U 1 4 6 Grn<br />

Guard milkvetch 2 1 1 0 0 1 U 5 6 Wsh; USFS: S<br />

Rollins’ sweetvetch<br />

Coal Cliffs sweetvetch<br />

2 1 1 0 0 U U 4 6 Uin<br />

2 1 1 0 0 U U 4 6 Crb, Dch, Emr;<br />

USFS: S<br />

Maybell locoweed 1 1 U 1 0 1 U 4 6 Dag<br />

Jones’ locoweed 2 1 U 1 0 0 U 4 6 Emr, Grf, Grn,<br />

Irn, Snp, Uin<br />

Aromatic breadroot<br />

Barneby’s breadroot<br />

1 1 U 1 0 1 1 5 6 Emr?, Grn<br />

1 1 1 1 0 1 U 5 6 Kan, Wsh; BLM:<br />

S<br />

Tuhy’s breadroot 2 1 0 1 0 U U 4 6 SnJ; BLM: S<br />

Pediomelum epipsilum Kane breadroot 2 1 0 1 0 1 U 5 6 Kan; BLM:S<br />

Pediomelum mephiticum Skunk breadroot 1 1 1 0 0 1 1 5 5 Wsh<br />

Pediomelum pariense Paria breadroot 2 1 1 1 0 0 U 5 6 Grf, Kan; USFS:<br />

S<br />

Pediomelum retrorsum<br />

(P. megalanthum var. r.)<br />

Psoralidium lanceolatum<br />

var. stenostachys<br />

Psorothamnus arborescens<br />

var. pubescens<br />

Psorothamnus nummularius<br />

(P. polydenius<br />

var. jonesii)<br />

Peach Springs<br />

breadroot<br />

Rydberg’s scurfpea<br />

Beauty indigobush<br />

1 1 U 1 0 1 U 4 6 Wsh<br />

2 1 U 1 0 0 U 4 6 Dav, Jub, Mil,<br />

Slt, Toe, Web<br />

1 1 1 1 0 U U 4 6 Kan<br />

Jones’ indigo-bush 2 1 0 1 0 U U 4 6 Emr; BLM: S


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Fabaceae<br />

(Leguminosae)<br />

Psorothamnus polydenius<br />

var. polydenius<br />

Glandular indigobush<br />

1 1 U 1 0 1 U 4 6 Wsh<br />

Trifolium beckwithii Beckwith’s clover 1 1 1 0 0 1 1 5 5 Piu?, Sev<br />

Gentianaceae<br />

Hydrangeaceae<br />

(Saxifragaceae)<br />

Swertia gypsicola<br />

(Frasera gypsicola)<br />

Jamesia americana var.<br />

zionis<br />

White River swertia<br />

1 1 1 1 0 1 U 5 6 Mil; BLM: S<br />

Zion jamesia 2 1 1 1 0 0 0 5 5 Kan, Wsh;<br />

USFS: S<br />

Jamesia tetrapetala Basin jamesia 1 1 1 1 0 U U 4 6 Mil; BLM: S;<br />

USFS: S<br />

Hydrophyllaceae Phacelia austromontana Southern phacelia 1 1 1 0 0 1 1 5 5 Wsh<br />

Phacelia cottamii Cottam’s phacelia 2 1 0 1 0 1 U 5 6 Crb, Emr, Sev<br />

Phacelia glandulosa<br />

Phacelia mammillarensis<br />

Glandular scorpion-weed<br />

Nipple Bench phacelia<br />

1 1 U 1 0 1 U 4 6 Grn, Uin<br />

2 1 0 1 0 1 U 5 6 Grf, Kan<br />

Phacelia palmeri Palmer’s phacelia 1 1 1 1 0 1 0 5 5 Wsh<br />

Phacelia perityloides<br />

var. laxiflora<br />

Crevice phacelia 1 1 1 1 0 1 0 5 5 Wsh<br />

Phacelia salina<br />

Phacelia splendens<br />

Bitter Creek scorpion-weed<br />

Eastwood’s phacelia<br />

1 1 1 1 0 U U 4 6 Snp, Toe<br />

1 1 U 1 0 1 U 4 6 Gra<br />

Iridaceae<br />

Juncaceae<br />

Lamiaceae<br />

(Labiatae)<br />

Liliaceae<br />

Phacelia tetramera<br />

Four-parted phacelia<br />

1 1 1 0 1 U U 4 6 Web<br />

Tricardia watsonii Three hearts 1 1 1 0 U 1 1 5 6 Wsh<br />

Sisyrinchium douglasii<br />

(Olsynium d.)<br />

Juncus tweedyi<br />

(J. brevicaudatus)<br />

Stachys rothrockii<br />

Allium geyeri var. chatterleyi<br />

Purple-eyed grass 1 1 1 U 0 1 U 4 6 Toe<br />

Tweedy’s rush 1 1 1 0 0 1 1 5 5 Box<br />

Rothrock’s hedgenettle<br />

1 1 U 1 0 1 U 4 6 Kan<br />

Chatterley’s onion 2 1 1 0 0 U U 4 6 Snj; USFS: S<br />

Allium passeyi Passey’s onion 2 1 0 1 0 U U 4 6 Box<br />

Loasaceae Eucnide urens Desert rock-nettle 1 1 1 1 0 1 U 5 6 Wsh<br />

Mentzelia goodrichii<br />

Goodrich’s<br />

stickleaf<br />

2 1 1 1 0 0 U 5 6 Dch; BLM: S;<br />

USFS: S<br />

225


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Loasaceae<br />

Malvaceae<br />

Najadaceae<br />

Mentzelia multicaulis<br />

var. flumensevera<br />

Mentzelia multicaulis<br />

var. uintahensis<br />

Sphaeralcea caespitosa<br />

var. caespitosa<br />

Najas caespitosa<br />

(included in N. flexilis<br />

by most authors)<br />

Sevier Canyon<br />

stickleaf<br />

Uinta Basin<br />

stickleaf<br />

Jones’ globemallow<br />

2 1 1 1 0 0 U 5 6 Piu, Sev<br />

1 1 U 1 0 1 U 4 6 Dch, Uin<br />

2 1 0 1 0 1 0 5 5 Bvr, Mil; BLM:<br />

S<br />

Fish Lake naiad 1 1 1 0 U 1 1 5 6 Sev; USFS: S<br />

Oleaceae Menodora spinescens Spiny menodora 1 1 1 1 0 1 0 5 5 Wsh<br />

Onagraceae Camissonia atwoodii Atwood’s camissonia<br />

2 1 0 1 0 1 U 5 6 Kan<br />

Camissonia bairdii Baird’s camissonia 1 1 1 1 0 1 U 5 6 Wsh; BLM: S<br />

Camissonia claviformis<br />

var. aurantiaca<br />

Camissonia claviformis<br />

var. claviformis<br />

Camissonia claviformis<br />

var. cruciformis<br />

Camissonia gouldlii<br />

Clubpod camissonia<br />

Clubpod camissonia<br />

Clubpod camissonia<br />

Gould’s camissonia<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

1 1 1 1 0 1 U 5 6 Mil, Wsh; BLM:<br />

S<br />

Epilobium nevadense<br />

Oenothera deltoides var.<br />

decumbens<br />

Ophioglossaceae Botrychium multifidum<br />

Orchidaceae<br />

Papaveraceae<br />

Habenaria zothecina<br />

(Platanthera z.)<br />

Eschscholzia mexicana<br />

(E. californica var. m.)<br />

Nevada willowherb<br />

St. George<br />

evening-primrose<br />

Leathery grape<br />

fern<br />

1 1 1 1 0 U U 4 6 Irn, Mil, Wsh;<br />

BLM: S; USFS:<br />

S<br />

1 1 U 1 0 1 U 4 6 Wsh<br />

1 1 1 0 1 U U 4 6 Dch<br />

Alcove bog-orchid 1 1 1 1 0 1 U 5 6 Emr, Grf, Grn,<br />

Snj, Uin<br />

Mexican goldenpoppy<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

Papaver coloradense<br />

(P. uintanense, P.<br />

kluanense, P. radicatum)<br />

Alpine Rocky<br />

Mountain poppy<br />

1 1 1 1 0 U U 4 6 Dag, Dch, Sum;<br />

USFS: S<br />

Platystemon californicus Creamcups 1 1 1 0 0 1 1 5 5 Wsh<br />

Poaceae<br />

(Gramineae)<br />

Andropogon glomeratus Bushy bluestem 1 1 1 1 0 U U 4 6 Grf, Kan, Snj,<br />

Way<br />

Festuca dasyclada <strong>Utah</strong> fescue 1 1 1 1 0 1 U 5 6 Emr, Grf, Snp<br />

226


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Poaceae<br />

(Gramineae)<br />

Polemoniaceae<br />

Imperata brevifolia Satintail 1 1 1 1 0 U U 4 6 Kan, Snj<br />

Panicum hallii Hall’s panicgrass 1 1 1 0 U U 1 4 6 Bvr<br />

Stipa arnowiae Arnow’s ricegrass 1 1 1 0 1 U U 4 6 Grf, Grn, Irn,<br />

Jub, Kan, Uin,<br />

Wsh<br />

Ipomopsis spicata var.<br />

spicata<br />

(Gilia s. var. s.)<br />

Ipomopsis tridactyla<br />

(I. spicata ssp. t., Gilia<br />

tridactyla)<br />

Phlox lutescens<br />

(P. austromontana var.<br />

lutescens)<br />

Spike gilia 1 1 1 1 0 U U 4 6 Dag<br />

Cedar Breaks gilia 2 1 1 1 0 0 U 5 6 Irn, Piu<br />

Yellowish phlox 2 1 1 0 0 U U 4 6 Grf, Grn, Snj<br />

Phlox opalensis Opal phlox 1 1 1 1 0 1 U 5 6 Dag<br />

Polygonaceae Eriogonum acaule Stemless wild<br />

buckwheat<br />

Eriogonum aretioides<br />

Eriogonum brevicaule<br />

var. loganum<br />

(E. loganum)<br />

Eriogonum cernuum var.<br />

psammophilum<br />

(var. not recognized by<br />

Holmgren et al. <strong>2012</strong>)<br />

Eriogonum corymbosum<br />

var. albiflorum<br />

(E. thompsoniae var. a.)<br />

Widtsoe wild<br />

buckwheat<br />

Logan wild buckwheat<br />

Sand Dune nodding<br />

wild buckwheat<br />

Virgin wild buckwheat<br />

1 1 1 1 0 U U 4 6 Rch<br />

2 1 1 1 0 0 U 5 6 Emr, Grf; USFS:<br />

S<br />

2 1 1 0 0 U U 4 6 Cch, Mor, Rch;<br />

USFS: S<br />

2 1 U 1 0 0 U 4 6 Grf, Kan, Snj;<br />

var. not recognized<br />

by Holmgren<br />

et al. (<strong>2012</strong>)<br />

1 1 U 1 0 1 U 4 6 Wsh<br />

Eriogonum corymbosum<br />

var. aureum (sensu<br />

stricto)<br />

Eriogonum ephedroides<br />

(E. brevicaule var.<br />

ephedroides)<br />

Eriogonum exaltatum<br />

(E. insigne)<br />

Eriogonum heermannii<br />

var. subspinosum<br />

Eriogonum mortonianum<br />

Eriogonum scabrellum<br />

Golden buckwheat 2 1 U 1 0 0 U 4 6 Wsh; does not<br />

include ‘var.<br />

glutinosum’<br />

Ephedra wild<br />

buckwheat<br />

Ladder wild buckwheat<br />

Tabeau Peak wild<br />

buckwheat<br />

Morton’s wild<br />

buckwheat<br />

Westwater wild<br />

buckwheat<br />

2 1 0 1 0 1 U 5 6 Uin<br />

1 1 1 1 0 1 0 5 5 Irn, Kan, Wsh<br />

1 1 1 1 0 1 0 5 5 Wsh<br />

2 1 1 0 0 U U 4 6 Kan<br />

1 1 1 1 0 U U 4 6 Emr, Grf, Grn,<br />

Kan, Snj<br />

227


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Polygonaceae<br />

Polypodiaceae<br />

Primulaceae<br />

Ranunculaceae<br />

Eriogonum wrightii var.<br />

wrightii<br />

Wright’s wild<br />

buckwheat<br />

1 1 U 1 0 1 U 4 6 Wsh<br />

Koenigia islandica Koenigia 1 1 1 1 0 U U 4 6 Dch<br />

Pterostegia drymarioides<br />

Adiantum pedatum var.<br />

aleuticum<br />

Pterostegia 1 1 U 1 0 1 U 4 6 Wsh<br />

Northern maidenhair<br />

fern<br />

1 1 1 1 1 0 U 5 6 Grf, Slt, Wsh<br />

Cheilanthes wootonii Wooton’s lip-fern 1 1 1 1 0 1 0 5 5 Wsh<br />

Cystopteris bulbifera<br />

Gymnocarpium dryopteris<br />

Dodecatheon pulchellum<br />

var. zionense<br />

(includes subvar. huberi)<br />

Bulblet bladder<br />

fern<br />

1 1 1 1 0 U U 4 6 Slt, Snj, Wsh<br />

Oak fern 1 1 1 1 0 U U 4 6 Piu<br />

Zion shooting-star 1 1 1 1 0 0 1 5 5 Crb, Grn, Kan,<br />

Snj?, Wsh<br />

Primula specuicola Cave primrose 1 1 1 1 0 1 U 5 6 Grf, Grn, Kan,<br />

Snj, Way<br />

Aquilegia atwoodii<br />

(included in A. fosteri by<br />

Holmgren et al. <strong>2012</strong>)<br />

Atwood’s columbine<br />

2 1 1 1 0 0 U 5 6 Uin; BLM: S<br />

Aquilegia barnebyi Shale columbine 1 1 1 1 0 1 U 5 6 Dch, Uin<br />

Aquilegia desolaticola<br />

Aquilegia fosteri<br />

(A. formosa var. fosteri,<br />

A. desertorum; may include<br />

A. atwoodii)<br />

Aquilegia grahamii<br />

(A. micrantha var. g.)<br />

Desolation Canyon<br />

columbine<br />

Foster’s columbine<br />

Graham’s columbine<br />

2 1 1 1 0 0 U 5 6 Grn; BLM: S<br />

2 1 1 1 0 0 0 5 5 Wsh<br />

2 1 1 1 0 0 U 5 6 Uin; USFS: S<br />

Rhamnaceae<br />

Rosaceae<br />

Aquilegia loriae<br />

(A. micrantha var. l.)<br />

Trautvetteria caroliniensis<br />

Ceanothus greggii var.<br />

franklinii<br />

Crataegus douglasii var.<br />

duchesnensis<br />

Lori’s columbine 2 1 1 1 0 0 0 5 5 Kan<br />

Carolina tassel-rue 1 1 1 1 0 U U 4 6 Snj<br />

Franklin’s desertlilac<br />

Duchesne hawthorn<br />

2 1 1 0 0 U U 4 6 Grf?, Grn, Snj<br />

2 1 1 0 0 U U 4 6 Dch, Uin, Was<br />

Ivesia utahensis <strong>Utah</strong> ivesia 2 1 1 0 0 1 U 5 6 Slt, Sum, Uta,<br />

Was; USFS: S<br />

Potentilla angelliae<br />

Angell’s cinquefoil<br />

228<br />

2 1 1 0 0 1 U 5 6 Way; USFS: S


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Rosaceae Rubus neomexicanus New Mexico<br />

thimbleberry<br />

Rutaceae<br />

Ptelea trifoliata var.<br />

lutescens<br />

1 1 1 1 0 U U 4 6 Grf, Snj<br />

Hoptree 1 1 1 1 0 U 1 5 6 Grf?, Kan, Wsh<br />

Salicaceae Salix arizonica Arizona willow 1 1 1 1 0 1 U 5 6 Grf, Irn, Snp,<br />

Sev; USFS: S<br />

Saururaceae Anemopsis californica Yerba mansa 1 1 1 1 0 U U 4 6 Uta, Wsh<br />

Scrophulariaceae<br />

Castilleja parvula var.<br />

parvula<br />

Maurandya antirrhiniflora<br />

Mimulus bigelovii var.<br />

cuspidatus<br />

Tushar paintbrush 2 1 1 1 0 0 U 5 6 Bvr, Grf, Piu;<br />

USFS: S<br />

Maurandya 1 1 1 0 0 1 1 5 5 Wsh<br />

Bigelow’s monkeyflower<br />

1 1 1 0 0 1 1 5 5 Wsh<br />

Mohavea breviflora Desert snapdragon 1 1 1 0 0 1 1 5 5 Wsh<br />

Penstemon abietinus Firleaf penstemon 2 1 1 0 0 U U 4 6 Sev, Uta<br />

Penstemon acaulis var.<br />

acaulis<br />

Penstemon ammophilus<br />

Penstemon angustifolius<br />

var. vernalensis<br />

Penstemon atwoodii<br />

Penstemon barbatus var.<br />

trichander<br />

Penstemon bracteatus<br />

Penstemon compactus<br />

Penstemon duchesnensis<br />

(P. dolius var. duchesnensis)<br />

Penstemon franklinii<br />

Stemless penstemon<br />

Sandloving penstemon<br />

2 1 0 1 0 1 U 5 6 Dag; BLM: S;<br />

USFS: S<br />

2 1 1 1 0 0 U 5 6 Grf, Kan, Wsh<br />

Vernal penstemon 2 1 1 0 0 1 U 5 6 Dag, Uin<br />

Atwood’s penstemon<br />

2 1 1 1 0 0 U 5 6 Grf, Kan<br />

Scarlet penstemon 1 1 1 1 0 U U 4 6 Snj<br />

Red Canyon penstemon<br />

Bear River penstemon<br />

Duchesne penstemon<br />

Franklin’s penstemon<br />

2 1 1 1 0 0 U 5 6 Grf; USFS: S<br />

2 1 1 1 0 0 U 5 6 Cch; USFS: S<br />

2 1 0 1 0 1 U 5 6 Dch<br />

2 1 1 0 0 1 U 5 6 Irn; BLM: S<br />

Penstemon idahoensis Idaho penstemon 1 1 1 1 0 U U 4 6 Box; BLM: S;<br />

USFS: S<br />

Penstemon marcusii<br />

Penstemon navajoa<br />

Marcus Jones’<br />

penstemon<br />

Navajo Mountain<br />

penstemon<br />

2 1 U 1 0 1 U 5 7 Crb, Emr<br />

2 1 1 0 0 1 U 5 6 Snj<br />

229


Pot Score<br />

Min Score<br />

Trend<br />

Threat<br />

Intrin Rar<br />

Hab Spec<br />

# Indiv<br />

# Pops<br />

Range<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 3. UNPS Rare <strong>Plant</strong> List: Watch List, continued<br />

Family Species Common Name<br />

County Dist.<br />

& Legal Status<br />

Scrophulariaceae<br />

Penstemon petiolatus<br />

Penstemon scariosus<br />

var. cyanomontanus<br />

Penstemon sepalulus<br />

Crevice penstemon<br />

Blue Mountain<br />

penstemon<br />

Littlecup penstemon<br />

1 1 1 1 0 1 0 5 5 Wsh<br />

2 1 1 1 0 0 U 5 6 Uin<br />

2 1 U 1 0 0 U 4 6 Jub, Uta, Wsh?<br />

Selaginellaceae Selaginella utahensis <strong>Utah</strong> spike-moss 2 1 1 1 0 0 0 5 5 Kan, Wsh<br />

Violaceae<br />

Viola frank-smithii<br />

Viola purpurea var.<br />

charlestonensis (V.<br />

charlestonensis)<br />

Bear River Range<br />

violet<br />

Charleston Mountain<br />

violet<br />

2 1 1 1 0 0 U 5 6 Cch; USFS: S<br />

1 1 1 1 0 1 0 5 5 Kan, Wsh;<br />

USFS: S<br />

Zygophyllaceae Fagonia laevis Fagonia 1 1 1 1 0 1 U 5 6 Wsh<br />

230


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List<br />

The following table includes 115 species with three or more ranking criteria scored as “unknown”. A large number<br />

of these species have only recently been named or discovered within <strong>Utah</strong>, and additional field surveys are needed to<br />

confirm their abundance, distribution, habitat needs, life history patterns, potential threats, and trends. Some species<br />

on the list have taxonomic questions that still need to be resolved. All of the plants included here have the potential<br />

to be ranked as extremely high or high priority, or as watch species, once needed studies are completed. Species are<br />

arranged alphabetically by family and species. Additional information is provided on county-level distribution (see<br />

Table 3 for codes) and data needs. Legal Status: Bureau of Land Management (BLM) and US Forest Service (USFS)<br />

Sensitive = S; US Fish and Wildlife Service (USFWS) Candidate = C.<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Apiaceae<br />

(Umbelliferae)<br />

Asteraceae<br />

(Compositae)<br />

Cymopterus basalticus<br />

Cymopterus crawfordensis<br />

Artemisia biennis var.<br />

diffusa<br />

Artemisia parryi<br />

Artemisia tridentata var.<br />

parishii<br />

Chrysothamnus nauseosus<br />

var. uintahensis<br />

(Ericameria x u.)<br />

Crepis runcinata var.<br />

aculeolata<br />

Shadscale springparsley<br />

Crawford Mountain<br />

spring-parsley<br />

Mystery wormwood<br />

Parry’s wormwood<br />

Parish’s big sagebrush<br />

Bvr, Mil<br />

Rch<br />

Grf<br />

Grn, Snj<br />

SW UT<br />

Info needed on # of indiviuals, threats, &<br />

trends<br />

Recently named, info needed on # of individuals,<br />

habitat specificity, threats, trends<br />

Taxonomic questions, info needed on # of<br />

populations, intrinsic rarity, threats, trends<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Info needed on distribution in UT, # of<br />

individuals, # of populations, threats, trends<br />

Uinta rabbitbrush Dag, Uin Info needed on # of individuals, threats, &<br />

trends<br />

<strong>Utah</strong> hawksbeard Kan Taxonomic questions, info needed on habitat<br />

specificity, threats, & trends<br />

Erigeron katiae Katie’s daisy Rch Newly described, info needed on habitat<br />

specificity, threats, & trends<br />

Erigeron mancus La Sal daisy Grn, Snj Recent research needs to be reviewed relating<br />

to # of populations, threats, & trends;<br />

USFS: S<br />

Erigeron watsonii Watson’s daisy Reported Info needed on # of individuals, habitat<br />

specificity, threats, & trends<br />

Haplopappus acaulis<br />

var. atwoodii<br />

(not recognized in FNA)<br />

Haplopappus crispus<br />

(Ericameria crispa)<br />

Haplopappus leverichii<br />

(Isocoma leverichii, I.<br />

humilis)<br />

Haplopappus racemosus<br />

var. paniculatus<br />

(Pyrrocoma racemosa<br />

var. paniculata)<br />

Haplopappus racemous<br />

var. prionophyllus<br />

(Pyrrocoma racemosa<br />

var. prionophylla)<br />

Atwood’s goldenweed<br />

Pine Valley<br />

goldenbush<br />

Canyon goldenweed<br />

Racemose goldenweed<br />

Racemose goldenweed<br />

Jub<br />

Mil?, Wsh;<br />

Wsh<br />

Mil<br />

Cch, Dch, Uta<br />

Treated as var. glabratus by Welsh et al.<br />

(2008); info needed on # of individuals,<br />

habitat specificity, threats, & trends<br />

Info needed on # of individuals, threats, &<br />

trends; BLM: S; USFS: S<br />

Taxonomic questions, info needed on intrinsic<br />

rarity, threats, & trends; not seen<br />

since 1971<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Info needed on # of individuals, threats, &<br />

trends<br />

231


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Asteraceae<br />

(Compositae)<br />

Brassicaceae<br />

(Cruciferae)<br />

Hofmeisteria pluriseta<br />

(Pleurocoronis p.)<br />

Lygodesmia grandiflora<br />

var. doloresensis<br />

(L. doloresensis)<br />

Arrowleaf Wsh? Info needed on # of individuals, threats, &<br />

trends<br />

Dolores River<br />

skeletonplant<br />

Grn?<br />

Confirmation needed whether this species<br />

is in UT; info needed on intrinsic rarity,<br />

threats, & trends; BLM: S<br />

Senecio bairdii Baird’s groundsel Box Newly described, info needed on habitat<br />

specificity, intrinsic rarity, threats, trends<br />

Senecio streptanthifolius<br />

var. platylobus<br />

Senecio werneriifolius<br />

var. malmstenoides<br />

Arabis goodrichii<br />

(Boechera g.)<br />

Arabis holboellii var.<br />

derensis<br />

(Boechera inyoensis,<br />

included in A. beckwithii<br />

by some authors)<br />

Arabis lasiocarpa<br />

(Boechera l.)<br />

Arabis perennans var.<br />

thorneae<br />

(Boechera selbyi var. t.)<br />

Arabis pulchra var.<br />

duchesnensis<br />

(Boechera duchesnensis)<br />

Arabis thompsonii<br />

(Boechera t., B. pallidifolia)<br />

Boechera glareosa<br />

(“Arabis glareosa”)<br />

Draba abajoensis<br />

(D. spectabilis var. glabrescens)<br />

Draba pedicellata var.<br />

pedicellata<br />

Wasatch groundsel Uta, Web<br />

Mt. Nebo groundsel<br />

Goodrich’s rockcress<br />

Desert Experimental<br />

Range rockcress<br />

Jub, Uta<br />

Mil<br />

Mil<br />

Wasatch rockcress Box, Cch, Rch,<br />

Slt, Uta<br />

Thorne’s rockcress Uin<br />

Duchesne rockcress<br />

Thompson’s rockcress<br />

Dch<br />

Snj<br />

Newly described, info needed on # of individuals,<br />

habitat specificity, threats, trends<br />

Newly described, info needed on # of individuals,<br />

threats, & trends<br />

Newly described, info needed on habitat<br />

specificity, threats, & trends; BLM: S<br />

Taxonomic questions, info needed on # of<br />

individuals, threats, & trends<br />

Info needed on habitat specificity, threats,<br />

& trends<br />

Recently described, info needed on habitat<br />

specificity, threats, & trends<br />

Taxonomic questions; info needed on # of<br />

individuals, habitat specificity, & trends<br />

Newly described, info needed on habitat<br />

specificity, threats, & trends<br />

Dorn’s rockcress Uin Recently described narrow endemic of CO<br />

& UT (holotype from S side of Blue Mountain),<br />

info needed on # of individuals, habitat<br />

specificity, number of populations,<br />

threats, & trends<br />

Abajo Peak draba Grn, Snj Recently described, info needed on # of<br />

individuals, intrinsic rarity, threats, &<br />

trends; USFS:S<br />

Cusick’s draba Toe Recently documented in UT, info needed<br />

on # of individuals, threats, and trends<br />

Draba pennellii Schell Creek draba Jub Recently documented in UT, info needed<br />

on # of individuals, threats, & trends;<br />

USFS: S<br />

Draba santaquinensis Santaquin draba Uta Recently described narrow endemic from<br />

southern Wasatch Range, info needed on #<br />

of individuals, habitat specificity, # of<br />

populations, threats, & trends; USFS: S<br />

232


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Brassicaceae<br />

(Cruciferae)<br />

Lepidium moabense<br />

(included in L. eastwoodiae<br />

by some authors)<br />

Physaria acutifolia var.<br />

repanda (included in P.<br />

grahamii in FNA)<br />

Physaria hemiphysaria<br />

var. hemiphysaria<br />

Moab pepperplant Grf, Grn, Kan, Snj Taxonomic questions, info needed on #<br />

of populations, intrinsic rarity, & trends<br />

Indian Canyon<br />

twinpod<br />

Skyline bladderpod<br />

Crb, Dch, Emr, Sev,<br />

Uin, Uta, Was<br />

Dch, Emr, Snp, Uta,<br />

Was<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Physaria hemiphysaria<br />

var. lucens<br />

Physaria navajoensis<br />

(Lesquerella navajoensis)<br />

Tavaputs bladderpod<br />

Crb<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Navajo bladderpod Kan? Taxonomic questions; info needed on #<br />

of individuals, intrinsic rarity, threats, &<br />

trends<br />

Physaria neeseae Neese’s twinpod Grf, Wsh? Newly described, info needed on # of<br />

individuals, threats, & trends<br />

Thelypodiopsis aurea Golden thelypody Snj Info needed on # of individuals, threats,<br />

& trends<br />

Thelypodiopsis vermicularis<br />

Wormwood thelypody<br />

Box, Irn, Jub, Mil,<br />

Snp, Sev, Toe, Uta<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Thelypodium rollinsii<br />

Rollins’ thelypody Bvr, Crb, Jub,<br />

Mil, Piu, Snp, Sev<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Cactaceae<br />

Coryphantha vivipara<br />

var. deserti<br />

Mohave pincushion<br />

cactus<br />

Irn, Wsh<br />

Info needed on # of individuals, number<br />

of populations, & trends<br />

Echinocactus polycephalus<br />

var. polycephalus<br />

Echinocactus polycephalus<br />

var. xeranthemoides<br />

(E. xeranthemoides)<br />

Opuntia pinkavae<br />

(O. basilaris var. woodburyi)<br />

Sclerocactus pubispinus<br />

var. pubispinus<br />

Manyhead barrel<br />

cactus<br />

Manyhead barrel<br />

cactus<br />

Pinkava’s pricklypear<br />

Great Basin fishhook<br />

Wsh?<br />

Kan?<br />

Kan, Wsh<br />

Bvr, Box, Irn, Jub,<br />

Mil, Toe<br />

Reports from UT need confirmation;<br />

info needed on # of individuals, # of<br />

populations, & trends<br />

Reports from UT need confirmation;<br />

info needed on # of individuals, # of<br />

populations, & trends<br />

May be represented by two forms in<br />

<strong>Utah</strong>; frequently hybridizes with other<br />

taxa; taxonomic problems; info needed<br />

on # of individuals, # of populations, &<br />

trends<br />

Info needed on # of individuals, # of<br />

populations, & trends<br />

Sclerocactus spinosior<br />

(S. pubispinus var. s.)<br />

Desert valley fishhook<br />

Jub, Mil, Sev<br />

Info needed on # of individuals, # of<br />

populations, & trends<br />

Caryophyllaceae Eremogone loisiae Lois’ sandwort Box, Cch, Dav, Jub,<br />

Rch, Slt, Snp, Toe,<br />

Uta, Web<br />

Recently named, info needed on # of<br />

individuals, threats, and trends<br />

233


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Chenopodiaceae Atriplex gardneri var.<br />

welshii (A. welshii)<br />

Fabaceae<br />

(Leguminosae)<br />

Atriplex powellii var.<br />

minuticarpa<br />

(A. minuticarpa)<br />

Astragalus brandegei<br />

Welsh’s saltbush Grn Taxonomic questions; info needed on #<br />

of individuals, threats, & trends<br />

Green River orach Emr, Gra, Way<br />

Brandegee’s milkvetch<br />

Emr, Grf, Irn, Piu,<br />

Sev, Way<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus callithrix<br />

Callaway milkvetch<br />

Mil<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus desperatus<br />

var. petrophilus<br />

Rock-loving milkvetch<br />

Emr<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus eastwoodiae<br />

Eastwood’s milkvetch<br />

Emr, Grf, Grn, Snj,<br />

Way<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus hornii Horn’s milkvetch Wsh? Reports for UT need confirmation; info<br />

needed on # of individuals, # of populations,<br />

& trends<br />

Astragalus kelseyae<br />

Astragalus laccoliticus<br />

(A. chamaeleuce var.<br />

laccoliticus)<br />

Astragalus lentiginosus<br />

var. negundo<br />

Astragalus lentiginosus<br />

var. stramineus<br />

Kelsey’s milkvetch<br />

Laccolite milkvetch<br />

Box Elder freckled<br />

milkvetch<br />

Web<br />

Grf, Way<br />

Box<br />

Newly described narrow endemic, info<br />

needed on # of individuals, habitat<br />

specificity, intrinsic rarity, & trends<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Newly described, info needed on # of<br />

individuals, threats, & trends<br />

Straw milkvetch Wsh? Reports from UT need confirmation;<br />

info needed on # of individuals, # of<br />

populations, & trends<br />

Astragalus pardalinus Panther milkvetch Emr, Grf, Grn, Way Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus pattersonii<br />

Patterson’s milkvetch<br />

Crb, Emr, Grf, Snj,<br />

Sev, Uin, Way<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus pinonis Pinyon milkvetch Bvr, Jub Info needed on habitat specificity,<br />

threats, & trends<br />

Astragalus preussii var.<br />

laxiflorus<br />

Littlefield milkvetch<br />

Wsh?<br />

Reports for UT need confirmation; info<br />

needed on # of individuals, # of populations,<br />

& trends<br />

Astragalus pubentissimus<br />

var. peabodianus<br />

Astragalus rafaelensis<br />

Peabody’s milkvetch<br />

San Rafael milkvetch<br />

Emr, Grn Taxonomic questions, info needed on #<br />

of individuals, threats, & trends; BLM:<br />

S<br />

Emr, Grn<br />

Info needed on # of individuals, threats,<br />

& trends<br />

Astragalus woodruffii<br />

Woodruff’s milkvetch<br />

Emr, Grf, Way<br />

Info needed on # of individuals, threats,<br />

& trends<br />

234


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Fabaceae<br />

(Leguminosae)<br />

Dalea flavescens var.<br />

epica (D. epica)<br />

Hole-in-the-Rock<br />

prairie-clover<br />

Grf, Snj<br />

Taxonomic questions; info needed on # of<br />

individuals, threats, & trends; BLM: S<br />

Lupinus flavoculatus Yellow-eye lupine Wsh Info needed on # of individuals, intrinsic<br />

rarity, & trends<br />

Pediomelum castoreum<br />

Trifolium andinum var.<br />

canone<br />

Beaver Dam<br />

breadroot<br />

Canyon Mountains<br />

clover<br />

Wsh?<br />

Mil<br />

Reports for UT need confirmation; info<br />

needed on # of individuals, # of populations,<br />

& trends<br />

Newly described, info needed on # of individuals,<br />

threats, & trends<br />

Trifolium andinum var.<br />

navajoense<br />

Trifolium andinum var.<br />

wahwahense<br />

Vicia americana var.<br />

lathyroides<br />

Navajo clover Snj Newly described, info needed on habitat<br />

specificity, threats, & trends<br />

Wah Wah clover Bvr Newly described, info needed on # of individuals,<br />

threats, & trends<br />

Pavant vetch Mil Newly described, info needed on habitat<br />

specificity, threats, & trends<br />

Gentianaceae Lomatogonium rotatum Marsh felwort Dag Info needed on # of individuals, threats, &<br />

trends<br />

Hydrangeaceae<br />

(Saxifragaceae)<br />

Jamesia americana var.<br />

rosea<br />

Hydrophyllaceae Phacelia crenulata var.<br />

orbicularis<br />

Rosy cliff jamesia Irn<br />

Henry Mountains<br />

phacelia<br />

Grf, Way<br />

Taxonomic questions, info needed on # of<br />

populations, intrinsic rarity, threats, trends<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Phacelia petrosa Forgotten phacelia Grf, Snj Info needed on # of individuals, threats, &<br />

trends<br />

Liliaceae Calochortus ciscoensis Cisco mariposa Dch, Grn, Uin Newly described, info needed on # of individuals,<br />

threats, & trends<br />

Loasaceae<br />

Mentzelia multicaulis<br />

var. librina<br />

Horse Canyon<br />

stickleaf<br />

Crb, Emr<br />

Info needed on # of individuals, threats, &<br />

trends; BLM: S<br />

Mentzelia thompsonii<br />

Thompson’s<br />

stickleaf<br />

Grn, Uin<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Petalonyx nitidus<br />

Shiny-leaf sandpaper-plant<br />

Wsh?<br />

Reports from UT need confirmation; info<br />

needed on # of individuals, # of populations,<br />

& trends<br />

Nyctaginaceae<br />

Abronia fragrans var.<br />

harrisii<br />

Harris’ fragrant<br />

sand-verbena<br />

Emr, Grf, Uin<br />

Taxonomic questions; info needed on # of<br />

populations, threats, & trends<br />

Onagraceae Camissonia bolanderi Bolander’s camissonia<br />

Emr, Way?<br />

Newly described, info needed on # of individuals,<br />

threats, & trends; BLM: S<br />

Ophioglossaceae Botrychium boreale<br />

(B. pinnatum)<br />

Northern grapefern<br />

Sum<br />

Taxonomic questions; info needed on habitat<br />

specificity, intrinsic rarity, threats, &<br />

trends<br />

Botrychium crenulatum Dainty moonwort Was Info needed on # of individuals, # of populations,<br />

threats, & trends; USFS: S<br />

235


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Ophioglossaceae Botrychium hesperium<br />

Papaveraceae<br />

Poaceae<br />

(Gramineae)<br />

Western moonwort<br />

Botrychium lanceolatum Lance-leaf grapefern<br />

Jub, Sum<br />

Jub<br />

Confirmation needed, info needed on # of<br />

individuals, # of populations, threats, &<br />

trends<br />

Info needed on habitat specificity, intrinsic<br />

rarity, threats, & trends<br />

Botrychium paradoxum Paradox moonwort Grf Confirmation needed; info needed on # of<br />

individuals, # of populations, threats, &<br />

trends; USFS: S<br />

Argemone corymbosa<br />

var. parva (A. parva)<br />

Bouteloua uniflora<br />

San Rafael<br />

prickly-poppy<br />

Grf, Grn, Snj,<br />

One-flower grama Reported, Zion<br />

NP<br />

Recently described; info needed on # of<br />

individuals, threats, & trends<br />

Confirmation needed; info needed on # of<br />

individuals, habitat specificity, # of populations,<br />

threats, & trends<br />

Leersia oryzoides Rice cutgrass Dav, Uta, Web Info needed on # of individuals, threats, &<br />

trends<br />

Stipa scribneri<br />

(Achnatherum s.)<br />

Scribner needlegrass<br />

Way<br />

Info needed on # of populations, threats, &<br />

trends<br />

Polemoniaceae<br />

Ipomopsis congesta var.<br />

goodrichii<br />

Goodrich gilia Dch Info needed on # of individuals, threats, &<br />

trends<br />

Langloisia schottii<br />

(Loeseliastrum s.)<br />

Navarretia furnissii<br />

Schott’s langloisia Wsh<br />

Furniss’s navarettia<br />

Cch, Sum, Was<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Recently named, info needed on # of individuals,<br />

intrinsic rarity, threats, & trends<br />

Phlox albomarginata<br />

White-margined<br />

phlox<br />

Rch<br />

Info needed on # of individuals, habitat<br />

specificity, threats, & trends<br />

Phlox austromontana<br />

var. jonesii<br />

(P. jonesii)<br />

Phlox austromontana<br />

var. prostrata<br />

Jones’ phlox Kan, Wsh Taxonomic questions; info needed on # of<br />

individuals, threats, & trends<br />

Silver Reef phlox Kan, Wsh Taxonomic questions; info needed on # of<br />

individuals, threats, & trends<br />

Polygonaceae<br />

Eriogonum brevicaule<br />

var. viridulum<br />

(E. viridulum)<br />

Duchesne wild<br />

buckwheat<br />

Dch, Uin<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Eriogonum contortum<br />

Grand Valley wild<br />

buckwheat<br />

Emr, Grn<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Eriogonum corymbosum<br />

var. hylophilum<br />

(E. hylophilum)<br />

Gate Canyon wild<br />

buckwheat<br />

Dch<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Eriogonum corymbosum<br />

var. nilesii<br />

Las Vegas wild<br />

buckwheat<br />

Kan?<br />

Authenticity of UT reports has been questioned<br />

by James Reveal (Holmgren et<br />

al.<strong>2012</strong>); info needed on # of individuals,<br />

habitat specificity, # of populations, &<br />

trends; BLM: S; USFWS: C<br />

236


Calochortiana <strong>December</strong> <strong>2012</strong> <strong>Number</strong> 1<br />

Appendix 4. UNPS Rare <strong>Plant</strong> List: Need Data List, continued<br />

Family Species Common Name County Dist. &<br />

Legal Status<br />

Information Needed<br />

Polygonaceae<br />

Rosaceae<br />

Scrophulariaceae<br />

Eriogonum corymbosum<br />

var. revealianum<br />

Eriogonum domitum<br />

Reveal’s wild<br />

buckwheat<br />

House Range wild<br />

buckwheat<br />

Eriogonum howellianum Howell’s wild<br />

buckwheat<br />

Eriogonum jamesii var.<br />

higginsii<br />

(E. arcuatum)<br />

Eriogonum lonchophyllum<br />

var. lonchophyllum<br />

Eriogonum microthecum<br />

var. tegetiforme<br />

(Lumped with var. lapidicola<br />

by Holmgren et<br />

al. <strong>2012</strong>)<br />

Eriogonum panguicense<br />

var. alpestre<br />

Eriogonum spathulatum<br />

var. kayeae<br />

(included in var. spathulatum<br />

by Holmgren et al.<br />

<strong>2012</strong>), Welsh et al.<br />

(2008) include E. artificis.<br />

Eriogonum spathulatum<br />

var. natum<br />

(E. natum)<br />

Potentilla diversifolia<br />

var. madsenii<br />

Penstemon acaulis var.<br />

yampaensis<br />

(P. yampaensis)<br />

Penstemon cyananthus<br />

var. judyae<br />

Higgins’ wild<br />

buckwheat<br />

Longleaf wild<br />

buckwheat<br />

Grf, Kan, Piu,<br />

Way<br />

Mil<br />

Jub, Mil, Toe<br />

Snj<br />

Emr, Grn, Snj,<br />

Uin<br />

Slender buckwheat Jub?, Mil, Wsh<br />

Cedar Breaks wild<br />

buckwheat<br />

Kaye’s wild buckwheat<br />

Son’s wild buckwheat<br />

Madsen’s cinquefoil<br />

Irn<br />

Bvr<br />

Mil<br />

Kan<br />

Yampa penstemon Dag<br />

Judy’s penstemon Uta<br />

var. heilii recently pulled out, updated<br />

status info needed on remaining pops, including<br />

# of individuals, threats, trends<br />

Described in 2011, endemic to House<br />

Range; info needed on # of individuals,<br />

habitat specificity, threats, & trends<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Variety not recognized by Reveal in Holmgren<br />

et al. (<strong>2012</strong>); info needed on # of individuals,<br />

threats, & trends<br />

Vars not recognized by Holmgren et al.<br />

(<strong>2012</strong>), including var. saurinum; info<br />

needed on # of individuals, threats, trends<br />

Taxonomic questions; info needed on # of<br />

individuals, threats, & trends<br />

Taxonomic questions; info needed on<br />

threats & trends (move to Watch list in<br />

future)<br />

Taxonomic questions. E. artificis considered<br />

a separate species by Reveal in Holmgren<br />

et al. (<strong>2012</strong>); info needed on # of individuals,<br />

threats, & trends; BLM: S<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Newly described, info needed on # of individuals,<br />

threats, & trends<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Recently described, info needed on # of<br />

individuals, threats, & trends<br />

Penstemon moffatii<br />

Mofatt penstemon Dch, Emr, Grf,<br />

Grn, Snj, Uta,<br />

Way<br />

Info needed on # of individuals, threats, &<br />

trends<br />

Penstemon nanus Dwarf penstemon Bvr, Irn?, Mil Info needed on # of individuals, threats, &<br />

trends<br />

237


<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

PO Box 520041<br />

Salt Lake City, UT 84152-0041<br />

The <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> was founded in 1978<br />

with a mission to promote the conservation, appreciation,<br />

and stewardship of native plants in the wild and in<br />

home cultivation. Through its publications, annual<br />

member meeting/dinner, field trips, and chapter meetings,<br />

UNPS is active in connecting citizens of <strong>Utah</strong> and<br />

the west with the native flora that makes the Beehive<br />

State so special. UNPS also funds an annual scholarship<br />

and small grants program using proceeds from its<br />

on-line store and generous contributions from members.<br />

Members of UNPS receive the <strong>Society</strong>’s bimonthly<br />

newsletter, the Sego Lily, discounts on posters, cds, and<br />

other merchandise at the UNPS store (www.unps.org),<br />

and are enrolled in their local chapter.<br />

Join hundreds of other <strong>Utah</strong> plant lovers (or plant<br />

lovers from other states or countries) in becoming a<br />

member of UNPS by submitting the form below or going<br />

online (www.unps.org).<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong> Membership<br />

__ New Member<br />

__ Renewal<br />

__ Gift Membership<br />

Membership Category<br />

__ Student $9.00<br />

__ Senior $12.00<br />

__ Individual $15.00<br />

__ Household $25.00<br />

__ Sustaining $40.00<br />

__ Supporting Organization $55.00<br />

__ Corporate $500.00<br />

__ Lifetime $250.00<br />

Mailing<br />

___ US Mail<br />

___ Electronic<br />

Contribution to UNPS scholarship fund ____ $<br />

Name _________________________________<br />

Street _________________________________<br />

City ______________________ State ________<br />

Zip ___________<br />

Email ___________________<br />

Chapter _______________________________<br />

__ Please send a complimentary copy of the Sego Lily to<br />

the above individual.<br />

Please enclose a check, payable to <strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

and send to:<br />

<strong>Utah</strong> <strong>Native</strong> <strong>Plant</strong> <strong>Society</strong><br />

PO Box 520041<br />

Salt Lake City, UT 84152-0041<br />

Join or renew on-line at unps.org<br />

238

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!