[Federal Register Volume 76, Number 36 (Wednesday, February 23, 2011)]
[Rules and Regulations]
[Pages 10166-10203]
From the Federal Register Online via the Government Publishing Office [www.gpo.gov]
[FR Doc No: 2011-3675]



[[Page 10165]]

Vol. 76

Wednesday,

No. 36

February 23, 2011

Part IV





Department of the Interior





-----------------------------------------------------------------------



Fish and Wildlife Service



-----------------------------------------------------------------------



50 CFR Part 17



Endangered and Threatened Wildlife and Plants; 12-Month Finding on a 
Petition To List Astragalus hamiltonii, Penstemon flowersii, Eriogonum 
soredium, Lepidium ostleri, and Trifolium friscanum as Endangered or 
Threatened; Rule

Federal Register / Vol. 76 , No. 36 / Wednesday, February 23, 2011 / 
Rules and Regulations

[[Page 10166]]


-----------------------------------------------------------------------

DEPARTMENT OF THE INTERIOR

Fish and Wildlife Service

50 CFR Part 17

Docket No. [FWS-R6-ES-2010-0087; MO 92210-0-008]


Endangered and Threatened Wildlife and Plants; 12-Month Finding 
on a Petition To List Astragalus hamiltonii, Penstemon flowersii, 
Eriogonum soredium, Lepidium ostleri, and Trifolium friscanum as 
Endangered or Threatened

AGENCY: Fish and Wildlife Service, Interior.

ACTION: Notice of 12-month petition finding.

-----------------------------------------------------------------------

SUMMARY: We, the U.S. Fish and Wildlife Service (Service), announce a 
12-month finding on a petition to list Astragalus hamiltonii (Hamilton 
milkvetch), Penstemon flowersii (Flowers penstemon), Eriogonum soredium 
(Frisco buckwheat), Lepidium ostleri (Ostler's peppergrass), and 
Trifolium friscanum (Frisco clover) as threatened or endangered under 
the Endangered Species Act of 1973 (ESA), as amended. After review of 
all available scientific and commercial information, we find that 
listing A. hamiltonii and P. flowersii is not warranted at this time. 
However, we ask the public to submit to us new information that becomes 
available concerning the threats to A. hamiltonii and P. flowersii or 
their habitat at any time. We find that listing E. soredium, L. 
ostleri, and T. friscanum as threatened or endangered is warranted. 
However, currently listing E. soredium, L. ostleri, and T. friscanum is 
precluded by higher priority actions to amend the Federal Lists of 
Endangered and Threatened Wildlife and Plants. Upon publication of this 
12-month petition finding, we will add E. soredium, L. ostleri, and T. 
friscanum to our candidate species list. We will develop proposed rules 
to list E. soredium, L. ostleri, and T. friscanum as our priorities 
allow. We will make determinations on critical habitat during 
development of the proposed listing rules. In the interim period, we 
will address the status of the candidate taxa through our annual 
Candidate Notice of Review.

DATES: The finding announced in this document was made on February 23, 
2011.

ADDRESSES: This finding is available on the Internet at http://www.regulations.gov at Docket Number FWS-R6-ES-2010-0087. Supporting 
documentation we used in preparing this finding is available for public 
inspection, by appointment, during normal business hours at the U.S. 
Fish and Wildlife Service, Utah Ecological Services Field Office, 2369 
West Orton Circle, Suite 50, West Valley City, UT 84119. Please submit 
any new information, materials, comments, or questions concerning this 
finding to the above address.

FOR FURTHER INFORMATION CONTACT: Larry Crist, Field Supervisor, U.S. 
Fish and Wildlife Service, Utah Ecological Services Field Office, 2369 
West Orton Circle, Suite 50, West Valley City, UT 84119; by telephone 
at 801-975-3330; or by facsimile at 801-975-3331mailto:. If you use a 
telecommunications device for the deaf (TDD), please call the Federal 
Information Relay Service (FIRS) at 800-877-8339.

SUPPLEMENTARY INFORMATION:

Background

    Section 4(b)(3)(B) of the ESA of 1973, as amended (16 U.S.C. 1531 
et seq.), requires that, for any petition to revise the Federal Lists 
of Endangered and Threatened Wildlife and Plants that contains 
substantial scientific or commercial information that listing a species 
may be warranted, we make a finding within 12 months of the date of 
receipt of the petition. In this finding, we will determine that the 
petitioned action is: (a) Not warranted, (b) warranted, or (c) 
warranted, but the immediate proposal of a regulation implementing the 
petitioned action is precluded by other pending proposals to determine 
whether species are threatened or endangered, and expeditious progress 
is being made to add or remove qualified species from the Federal Lists 
of Endangered and Threatened Wildlife and Plants. Section 4(b)(3)(C) of 
the ESA requires that we treat a petition for which the requested 
action is found to be warranted but precluded as though resubmitted on 
the date of such finding, that is, requiring a subsequent finding to be 
made within 12 months. We must publish these 12-month findings in the 
Federal Register.

Previous Federal Actions

    On July 30, 2007, we received a petition dated July 24, 2007, from 
Forest Guardians (now WildEarth Guardians), requesting that the 
Service: (1) Consider all full species in our Mountain Prairie Region 
ranked as G1 or G1G2 by the organization NatureServe, except those that 
are currently listed, proposed for listing, or candidates for listing; 
and (2) list each species as either endangered or threatened. The 
petition included the five plant species addressed in this finding. The 
petition incorporated all analysis, references, and documentation 
provided by NatureServe in its online database at http://www.natureserve.org/. The document clearly identified itself as a 
petition and included the petitioners' identification information, as 
required in 50 CFR 424.14(a). We sent a letter to the petitioners, 
dated August 24, 2007, acknowledging receipt of the petition and 
stating that, based on preliminary review, we found no compelling 
evidence to support an emergency listing for any of the species covered 
by the petition.
    On March 19, 2008, WildEarth Guardians filed a complaint (1:08-CV-
472-CKK) indicating that the Service failed to comply with its 
mandatory duty to make a preliminary 90-day finding on their two 
multiple species petitions--one for mountain-prairie species and one 
for southwest species.
    On June 18, 2008, we received a petition from WildEarth Guardians, 
dated June 12, 2008, to emergency list 32 species under the 
Administrative Procedure Act and the ESA. Of those 32 species, 11 were 
included in the July 24, 2007, petition to be listed on a nonemergency 
basis. Although the ESA does not provide for a petition process for an 
interested person to seek to have a species emergency listed, section 
4(b)(7) of the ESA authorizes the Service to issue emergency 
regulations to temporarily list a species. In a letter dated July 25, 
2008, we stated that the information provided in both the 2007 and 2008 
petitions and in our files did not indicate that an emergency situation 
existed for any of the 11 species.
    On February 5, 2009 (74 FR 6122), we published a 90-day finding on 
165 species from the petition to list 206 species in the mountain-
prairie region of the United States as endangered or threatened under 
the ESA. We found that the petition did not present substantial 
scientific or commercial information indicating that listing was 
warranted for these species and, therefore, did not initiate further 
status reviews in response to the petition. Two additional species were 
reviewed in a concurrent 90-day finding and again, we found that the 
petition did not present substantial scientific or commercial 
information indicating that listing was warranted for these species. 
Therefore we did not consider these two species further. For the 
remaining 39 species, we deferred our findings until a later date. One 
species of the 39 remaining species, Sphaeralcea gierischii (Gierisch

[[Page 10167]]

mallow), was already a candidate species for listing; therefore, 38 
species remained for consideration. On March 13, 2009, the Service and 
WildEarth Guardians filed a stipulated settlement in the District of 
Columbia Court, agreeing that the Service would submit to the Federal 
Register a finding as to whether WildEarth Guardians' petition 
presented substantial information indicating that the petitioned action 
may be warranted for 38 mountain-prairie species by August 9, 2009 
(WildEarth Guardians vs. Salazar 2009, case 1:08-CV-472-CKK).
    On August 18, 2009, we published a notice of 90-day finding (74 FR 
41649) on 38 species from the petition to list 206 species in the 
mountain-prairie region of the United States as endangered or 
threatened under the ESA. Of the 38 species, we found that the petition 
presented substantial scientific and commercial information for 29 
species, indicating that listing may be warranted for those 29 species. 
The 5 species we address in this 12-month finding were included in 
these 29 species. We initiated a status review of the 29 species to 
determine if listing was warranted. We also opened a 60-day public 
comment period to allow all interested parties an opportunity to 
provide information on the status of the 29 species. The public comment 
period closed on October 19, 2009. We received 224 public comments. Of 
these, two specifically addressed Astragalus hamiltonii, Penstemon 
flowersii, Eriogonum soredium, Lepidium ostleri, and Trifolium 
friscanum. All information received has been carefully considered in 
this finding. This notice constitutes the 12-month finding on the July 
24, 2007, petition to list five species (A. hamiltonii, P. flowersii, 
E. soredium, L. ostleri, and T. friscanum) as endangered or threatened.

Species Information--Astragalus hamiltonii

Taxonomy and Species Description
    Astragalus hamiltonii is a bushy perennial plant in the bean family 
(Fabaceae) that can grow up to 24 inches (in) (60 centimeters (cm)) 
tall (Welsh et al. 2003, p. 374). It has several sparsely leafed stems, 
with three to five (sometimes seven) leaflets per leaf, each 0.8 to 1.6 
in (2 to 4 cm) long and 0.2 to 0.4 in (5 to 10 millimeters (mm)) wide 
(Heil and Melton 1995a, p. 6). The terminal leaflet (at the tip of the 
leaf) is typically the largest leaflet (NatureServe 2009a, p. 3). In 
May and June, a single A. hamiltonii plant will produce many flowering 
stalks, with each stalk bearing 7 to 30 cream-colored flowers (Welsh et 
al. 2003, p. 374; NatureServe 2009a, p. 3). The fruits are hanging pods 
and usually mature by the end of June (NatureServe 2009a, p. 3).
    Astragalus hamiltonii was first described in 1952 (Porter 1952, pp. 
159-160). Although it was once considered a variety of A. lonchocarpus 
(Isely 1983, p. 422), A. hamiltonii is currently accepted as a distinct 
species, based on leaflet characteristics and geographic segregation 
(Barneby 1989, p. 72; Welsh et al. 2003, p. 374).
Distribution and Population Status
    Astragalus hamiltonii occurs generally west and southwest of 
Vernal, Utah. The species is found on Bureau of Land Management (BLM) 
land, the Uintah and Ouray Indian Reservation (hereafter ``Tribal'') 
lands, State of Utah School and Institutional Trust Lands 
Administration (SITLA) lands, and private lands across an approximate 
area 10 mile (mi) (16.1 kilometer (km)) by 20 mi (32.2 km) (Figure 1). 
We do not have comprehensive survey information for A. hamiltonii. 
Therefore, we do not know the full extent of the species' distribution 
or if the distribution has changed over time.
    The Utah Natural Heritage Program (UNHP) designates 11 element 
occurrences for Astragalus hamiltonii (UNHP 2010a, entire). Element 
occurrences are the specific locations, or sites, where plants are 
documented. Distinct element occurrences are identified if there is 
either 0.6 mi (1 km) of unsuitable habitat or 1.2 mi (2 km) of 
unoccupied, suitable habitat separating them (NatureServe 2004, p. 14).
    Astragalus hamiltonii element occurrences are based on collections 
of herbarium specimens. Two of the element occurrences identified by 
the UNHP were from Colorado and the southeast corner of the Uinta 
Basin, but we believe these locations are likely A. lonchocarpus, based 
on leaf characteristics and geographic distribution (NatureServe 2009a, 
p. 1; Goodrich 2010a, entire), so they are not considered further in 
this finding. Hereafter, we base our analysis on the remaining nine 
element occurrences (Table 1; Goodrich 2010b, entire).
    To determine the currently known distribution of Astragalus 
hamiltonii, we mapped the nine UNHP element occurrences (Figure 1). The 
UNHP records element occurrences using the public land survey system to 
the nearest quarter-quarter of the township, range, and section (UNHP 
2010a, entire). These element occurrences were the basis for our 
``population areas,'' but the population areas' boundaries were 
expanded to the nearest quarter-quarter of the township, range, and 
section, to encompass the location data from the 2010 surveys (Table 1; 
Goodrich 2010b, entire). This mapping approach resulted in some of the 
newly created population areas' perimeters eventually abutting adjacent 
population areas (Table 1; Figure 1). Large areas of potential habitat 
remain unsurveyed, so it is possible that the species is continuous 
across its range, or occurs outside of our identified population areas 
(Figure 1).
BILLING CODE 4310-55-P

[[Page 10168]]

[GRAPHIC] [TIFF OMITTED] TR23FE11.000

BILLING CODE 4310-55-C

[[Page 10169]]



                                              Table 1--Astragalus hamiltonii Plants Counted in 2010 Surveys
--------------------------------------------------------------------------------------------------------------------------------------------------------
                                                               Percent land ownership
                Population area                 ----------------------------------------------------        Number of Astragalus hamiltonii plants
                                                     BLM         SITLA        Tribal      Private
--------------------------------------------------------------------------------------------------------------------------------------------------------
1..............................................           11           54            0           35  Not counted.
2..............................................           76           13            1           11  4,863.
3..............................................           44           56            0            0  544.
4..............................................            0            0           10           90  15.
5..............................................            0            0           89           11  60.
6..............................................           57            5            0           38  10.
7..............................................            0            0           52           48  345.
8..............................................           13           62            0           25  Not counted.
9..............................................            0            0           81           19  Not counted.
                                                --------------------------------------------------------------------------------------------------------
    Total......................................           30           18           23           28  5,837.
--------------------------------------------------------------------------------------------------------------------------------------------------------

    We do not have long-term population count or trend information. The 
total population of Astragalus hamiltonii was estimated at 10,000 to 
15,000 plants in 1995 (Heil and Melton 1995a, p. 13). However, we do 
not know how this estimate was derived. In 2010, the U.S. Forest 
Service (USFS) counted over 5,800 A. hamiltonii individuals on BLM 
lands in areas west of Vernal in the vicinity of six of the element 
occurrences (numbers 2 to 7) (Table 1; Goodrich 2010b, entire). These 
were partial surveys that included revisits to six element occurrences.
    Astragalus hamiltonii is distributed sparsely across the landscape 
at low densities, but in optimum habitat A. hamiltonii can grow at 
densities of one to two plants per square yard (yd\2\) (square meter 
(m\2\)) (Heil and Melton 1995a, p. 13). Because A. hamiltonii is 
scattered across the landscape with unsurveyed, potential habitat 
between known sites, we believe the known element occurrences may be 
linked by contiguous habitat, and may either be one large population or 
a series of populations within a metapopulation.
Habitat
    Astragalus hamiltonii is a narrow endemic that grows on soils of 
the Duchesne River formation (Heil and Melton 1995a, p. 10; Goodrich 
2010c, pp. 13, 15). Less frequently, it is found in Mowry Shale and 
Dakota formations (Welsh et al. 2003, p. 374). A. hamiltonii is 
typically found on benches and steep slopes at elevations of 4,900 to 
6,200 feet (ft) (1,500 to 1,900 meters (m)). A. hamiltonii grows in 
red, erosive, sandy clay loam soils (Heil and Melton 1995a, pp. 10, 16; 
NatureServe 2009a, p. 3; Brunson 2010a, p. 1), and is associated with 
low-density desert shrub and juniper communities (Goodrich et al. 1999, 
p. 263; NatureServe 2009a, p. 3).
    Astragalus hamiltonii grows in old road cuts and road beds, 
sometimes quite robustly and producing abundant flowers and fruit 
(Goodrich et al. 1999, p. 263). Therefore, we believe the species may 
be able to tolerate moderate soil disturbances (Neese and Smith 1982, 
p. 36; Goodrich et al. 1999, p. 263).
Life History
    Astragalus hamiltonii growth, seedling establishment, and juvenile 
mortality are probably correlated with rainfall (Heil and Melton 1995a, 
p. 14). We do not know the reproductive system for this species, but it 
is assumed to reproduce mainly by outcrossing (cross-fertilization) 
(Heil and Melton 1995a, p. 14). Plants that are obligate outcrossers 
are self-incompatible, meaning they cannot fertilize themselves and, 
therefore, rely on other individuals of differing genetic make-up to 
reproduce (Stebbins 1970, p. 310).

Summary of Information Pertaining to the Five Factors--Astragalus 
hamiltonii

    Section 4 of the ESA (16 U.S.C. 1533) and implementing regulations 
(50 CFR part 424) set forth procedures for adding species to the 
Federal Lists of Endangered and Threatened Wildlife and Plants. Under 
section 4(a)(1) of the ESA, a species may be determined to be 
endangered or threatened based on any of the following five factors:
    (A) The present or threatened destruction, modification, or 
curtailment of its habitat or range;
    (B) Overutilization for commercial, recreational, scientific, or 
educational purposes;
    (C) Disease or predation;
    (D) The inadequacy of existing regulatory mechanisms; or
    (E) Other natural or manmade factors affecting its continued 
existence.
    In making our 12-month finding on the petition, we considered and 
evaluated the best available scientific and commercial information 
pertaining to Astragalus hamiltonii for the five factors provided in 
section 4(a)(1) of the ESA.
    In considering what factors might constitute threats to a species, 
we must look beyond the exposure of the species to a particular factor 
to evaluate whether the species may respond to that factor in a way 
that causes actual impacts to the species. If there is exposure to a 
factor and the species responds negatively, the factor may be a threat 
and, during the status review, we attempt to determine how significant 
a threat it is. The threat is significant if it drives, or contributes 
to, the risk of extinction of the species such that the species 
warrants listing as endangered or threatened as those terms are defined 
in the ESA. However, the identification of factors that could impact a 
species negatively may not be sufficient to compel a finding that the 
species warrants listing. The information must include evidence 
sufficient to suggest that these factors are operative threats that act 
on the species to the point that the species may meet the definition of 
endangered or threatened under the ESA.

Factor A. The Present or Threatened Destruction, Modification, or 
Curtailment of Its Habitat or Range

    The following factors may affect the habitat or range of Astragalus 
hamiltonii: (1) Conversion to agricultural use, (2) livestock grazing, 
(3) recreational activities, (4) oil and gas exploration and 
development, (5) nonnative invasive species, and (6) tar sands 
extraction.
(1) Conversion to Agricultural Use
    Astragalus hamiltonii grows on private and Tribal lands that can be 
used for agriculture. Agricultural land conversion is a change in land 
use to an agricultural use, including crops and pastures. The 
conversion to agricultural use results in the loss and fragmentation

[[Page 10170]]

of native plant habitats, including habitats of A. hamiltonii.
    Conversion of natural lands to agriculture historically impacted 
populations of Astragalus hamiltonii (Heil and Melton 1995a, p. 16), 
particularly in the four population areas where land ownership is 
private or Tribal. However, most of this development was limited to 
lower-lying areas outside of A. hamiltonii habitat (National 
Agriculture Imagery Program (NAIP) 2009, entire). It is likely that 
most of the suitable land in Uintah County, where irrigation water was 
available, was converted to agricultural use by 1970 (Hilton 2010, p. 
1). Major changes in the amount of agricultural land in Uintah County 
are not expected in the future (Hilton 2010, p. 2). Although historical 
conversion to agricultural use may have negatively impacted A. 
hamiltonii, we have no evidence to indicate that this factor is a 
threat to this species now or for the foreseeable future.
(2) Livestock Grazing
    Livestock grazing may result in the direct loss or damage to plants 
and their habitat through trampling, soil compaction, increased 
erosion, invasion of noxious weeds, and disturbance to pollinators 
(Kauffman et al. 1983, p. 684; Fleischner 1994, entire; Kearns et al. 
1998, p. 90; DiTomaso 2000, p. 257). All BLM lands where Astragalus 
hamiltonii is documented are within grazing allotments, including 
portions of population areas 1, 2, 3, 6, and 8 (see Table 1). In 2010, 
of all A. hamiltonii counted, 5,417 individuals (93 percent) occur in 
existing grazing allotments. We have no information on the extent of 
grazing on private or Tribal lands.
    We do not have any information concerning how grazing may affect 
this species. However, cattle tend to spend more time on gentle slopes 
(Van Buren 1982 in Fleischner 1994, p. 637). Astragalus hamiltonii 
grows on steep, erosive hillsides, and we believe this habitat 
preference offers some protection from livestock grazing and trampling. 
In addition, the grazing allotments that overlap A. hamiltonii sites on 
BLM land are fall and winter allotments (BLM 2008a, Appendix J); thus, 
A. hamiltonii is not actively growing or palatable when livestock are 
grazing these areas.
    In summary, the species occurs in areas that are subject to 
livestock grazing. However, the fall-winter season of grazing greatly 
reduces the chance that the plants are eaten by livestock. Astragalus 
hamiltonii typically grows on steep slopes and can occur on disturbed 
soils, which minimizes negative effects from livestock trampling within 
A. hamiltonii habitat. Therefore, we do not believe that livestock 
grazing is a threat to A. hamiltonii now or for the foreseeable future.
(3) Recreational Activities
    Off-highway vehicle (OHV) and recreational trail use (e.g., 
mountain bikes and motorized bikes) may result in direct loss or damage 
to plants and their habitat through soil compaction, increased erosion, 
invasion of noxious weeds, and disturbance to pollinators and their 
habitat (Eckert et al. 1979, entire; Lovich and Bainbridge 1999, p. 
316; Ouren et al. 2007, entire; BLM 2008a, pp. 4-94; Wilson et al. 
2009, p. 1).
    The OHV and recreational trail use occurs across the landscape 
where Astragalus hamiltonii grows. The OHV use is largely limited to 
existing roads and trails on BLM lands, which account for approximately 
a third of A. hamiltonii's known range (Table 1) (BLM 2008b, p. 46). 
There are no OHV restrictions on private or Tribal lands, but the 
species' association with steep, erosive hillsides likely minimizes OHV 
use in the species' habitat.
    Unauthorized off-road use occurs in Astragalus hamiltonii habitat 
in population area 2 (Brunson 2010a, p. 3). However, we observed plants 
growing directly next to these recreational trails (Brunson 2010a, p. 
3). As previously described, A. hamiltonii grows along road cuts and 
other disturbed areas, suggesting it can persist with some level of 
disturbance. We do not believe that the observed unauthorized off-road 
use is negatively impacting A. hamiltonii.
    In summary, the species' habitat preference for steep slopes, its 
ability to grow in disturbed soils, and off-road restrictions on BLM 
lands minimize the impacts of recreational use to Astragalus 
hamiltonii. Thus, we do not believe that recreational activities are a 
threat to A. hamiltonii now or for the foreseeable future.
(4) Oil and Gas Exploration and Development
    The effects of oil and gas exploration and development include 
increased vehicle traffic and removal of soil and vegetation when 
wells, roads, and associated infrastructure are built (BLM 2008c, pp. 
448-449). These disturbances can affect rare plant species through 
habitat destruction, habitat fragmentation, soil disturbance, spread of 
invasive weeds, and production of fugitive dust (particulate matter 
suspended in the air by wind and human activities) (BLM 2008c, pp. 448-
449).
    Energy exploration and development occurs across Astragalus 
hamiltonii's known range, but only in localized areas with small 
numbers of wells (Utah Division of Oil, Gas, and Mining (UDOGM) 2010, 
p. 1). Only one well is producing in A. hamiltonii habitat, and another 
well is currently being drilled. Seventeen wells were plugged and 
abandoned, most prior to 1976 (Gordon 2010a, pers. comm.; UDOGM 2010, 
p. 1). Plugged and abandoned wells are no longer in use and are usually 
recontoured and revegetated to match the surrounding landscape (Gordon 
2010b, pers. comm.). Plugged and abandoned wells also do not receive 
regular truck traffic like producing wells, so fugitive dust is less of 
an issue (Gordon 2010b, pers. comm.). Occasionally, plugged and 
abandoned wells may be reopened, disturbing areas that were previously 
reclaimed. If all the plugged and abandoned wells in A. hamiltonii 
habitat were reopened, this is still a small number of wells throughout 
the species' range.
    Large portions of population areas 1, 2, 3, 6, 7, and 8 (Table 1) 
are overlapped by oil and gas leases on state, Tribal, and BLM land. 
Two BLM oil and gas leases in population area 2 overlap more than 4,000 
known Astragalus hamiltonii individuals (UDOGM, 2010, p. 2). However, 
no oil or gas is being produced under these leases (UDOGM 2010, p. 2).
    The lack of oil and gas development in Astragalus hamiltonii 
habitat is most likely because there is not enough of those products 
currently obtainable to be economically feasible using current 
extraction technology (Doyle 2010, pers. comm.; Sparger 2010, pers. 
comm.) rendering dense energy developments unlikely in this area for 
the next 20 years (BLM 2008c, p. 486). Although some oil and gas 
development may occur in A. hamiltonii habitat, we would not expect it 
at densities that would significantly impact the species. Furthermore, 
A. hamiltonii is adapted to at least some disturbance and may be 
afforded additional protection by its tendency to grow on steep slopes 
that may be unsuitable for energy development. Therefore, oil and gas 
development is unlikely to occur in the foreseeable future at densities 
that would significantly impact the species.
    In summary, there is little oil and gas development within 
Astragalus hamiltonii habitat. Based on current technologies and low 
economic feasibility, we do not anticipate substantial development in 
the foreseeable future that would meaningfully impact the species. 
Therefore, we do not believe that oil and

[[Page 10171]]

gas exploration and development is a threat to A. hamiltonii now or in 
the foreseeable future.
(5) Nonnative Invasive Species
    The spread of nonnative invasive species is considered the second 
largest threat to imperiled plants in the United States (Wilcove et al. 
1998, p. 608). Invasive plants--specifically exotic annuals--negatively 
affect native vegetation, including rare plants. One of the most 
substantial effects is the change in vegetation fuel properties that, 
in turn, alter fire frequency, intensity, extent, type, and seasonality 
(Menakis et al. 2003, pp. 282-283; Brooks et al. 2004, p. 677; McKenzie 
et al. 2004, p. 898). Shortened fire return intervals make it difficult 
for native plants to reestablish or compete with invasive plants 
(D'Antonio and Vitousek 1992, p. 73).
    Invasive plants can exclude native plants and alter pollinator 
behaviors (D'Antonio and Vitousek 1992, pp. 74-75; DiTomaso 2000, p. 
257; Mooney and Cleland 2001, p. 5449; Levine et al. 2003, p. 776; 
Traveset and Richardson 2006, pp. 211-213). For example, Bromus 
tectorum outcompetes native species for soil nutrients and water 
(Melgoza et al. 1990, pp. 9-10; Aguirre and Johnson 1991, pp. 352-353).
    Bromus tectorum (cheatgrass) is a particularly problematic 
nonnative invasive annual grass in the Intermountain West. If already 
present in the vegetative community, B. tectorum increases in abundance 
after a wildfire, increasing the chance for more frequent fires 
(D'Antonio and Vitousek 1992, pp. 74-75). In addition, B. tectorum 
invades areas in response to surface disturbances (Hobbs 1989, pp. 389, 
393, 395, 398; Rejmanek 1989, pp. 381-383; Hobbs and Huenneke 1992, pp. 
324-325, 329, 330; Evans et al. 2001, p. 1308). B. tectorum is likely 
to increase due to climate change (see Factor E) because invasive 
annuals increase biomass and seed production at elevated levels of 
carbon dioxide (Mayeux et al. 1994, p. 98; Smith et al. 2000, pp. 80-
81; Ziska et al. 2005, p. 1328).
    Bromus tectorum occurs in Astragalus hamiltonii habitat (Brunson 
2010a, p. 1). However, B. tectorum and other invasive species are 
uncommon in many of the erosive red soils that A. hamiltonii prefers 
(Brunson 2010a, p. 1; Goodrich 2010c, p. 59). We do not anticipate a 
high degree of surface disturbances in A. hamiltonii habitats in the 
foreseeable future from other factors, such as livestock grazing or oil 
and gas development (Factor A).
    In summary, we know that invasive species can impact plant 
communities by increasing fire frequencies, outcompeting native 
species, and altering pollinator behaviors. These factors could be 
exacerbated by climate change patterns. However, invasive species do 
not occur in high densities in Astragalus hamiltonii habitat. Based on 
this fact and the limited amount of surface-disturbing activities 
within the species' habitat, we do not anticipate that nonnative 
invasive species densities will increase significantly, even with 
climate change. Therefore, we do not believe nonnative invasive 
species, or associated fires, are a threat to A. hamiltonii now or for 
the foreseeable future.
(6) Tar Sands Extraction
    The Duchesne River Formation, where most known Astragalus 
hamiltonii individuals occur, would be one of the formations targeted 
by tar sands extraction (BLM 2008d, p. 9). Tar sands extraction 
disturbs the soil surface and removes existing vegetation (BLM 2008d, 
p. 27). Impacts are similar to those described above in the Oil and Gas 
Exploration and Development section. Tar sands mining could result in 
the loss of A. hamiltonii individuals and their habitats.
    Tar sands leases are proposed for sale on BLM and State Lands along 
Asphalt Ridge southwest of Vernal, Utah (UDOGM 2010, p. 3). These lease 
parcels do not overlap known Astragalus hamiltonii sites, but they 
overlap with unsurveyed potential habitat within portions of population 
area 1.
    Tar sands leases are still in the proposal phase and there are 
currently no commercial tar sands operations on public lands in Utah 
(BLM 2008d, p. 4). High production costs and environmental issues are 
barriers to tar sands development in the United States (Bartis et al. 
2005, pp. 15, 53; Engemann and Owyang 2010, entire). Tar sands 
extraction may be feasible if the cost of crude oil becomes high enough 
in the future, but these high price projections are not expected to be 
realized until at least 2030 (Engemann and Owyang 2010, p. 2), and even 
then the environmental issues will need to be resolved.
    In summary, tar sands leases do not overlap a majority of 
Astragalus hamiltonii habitat. Large-scale, commercially viable 
development is not anticipated in the foreseeable future. Therefore, 
tar sands development is not considered a threat to A. hamiltonii now 
or in the foreseeable future.
Summary of Factor A
    Based on the best available information, we have concluded that 
conversion to agricultural use, livestock grazing, recreational 
activities, nonnative invasive species, oil and gas exploration and 
development, or tar sands extraction do not threaten Astragalus 
hamiltonii now or in the foreseeable future. Conversion to agricultural 
use probably resulted in historical loss of some A. hamiltonii habitat, 
but we do not anticipate ongoing conversions to agricultural use in the 
future. In addition, most agricultural use occurs in low-lying areas 
outside of the species' distribution. A. hamiltonii is protected from 
livestock grazing due to its habitat preference for steep hillsides and 
the fall-winter grazing season of the associated allotments. 
Recreational use is not a threat to A. hamiltonii because BLM restricts 
off-trail use. Where off-trail use occurs on private, State, and Tribal 
lands, the adaptation of A. hamiltonii to steep slopes and disturbed 
soils allows it to persist with moderate habitat disturbance. A. 
hamiltonii soils do not appear to support invasive plant species at 
densities needed to sustain wildfires. We also do not anticipate 
increased surface disturbances that could encourage the establishment 
of invasive species in A. hamiltonii habitat. Although energy 
development leases overlap A. hamiltonii habitat, it is unlikely that 
current technologies and economic conditions will support oil and gas 
or tar sands development in this area in the foreseeable future. Thus, 
the present or threatened destruction, modification, or curtailment of 
the habitat or range is not a threat to A. hamiltonii now or in the 
foreseeable future.

Factor B. Overutilization for Commercial, Recreational, Scientific, or 
Educational Purposes

    Astragalus hamiltonii is not a plant of horticultural interest. We 
are not aware of any instances where A. hamiltonii was collected from 
the wild other than as voucher specimens to document occurrences (UNHP 
2010a, entire). Therefore, we do not consider overutilization a threat 
to the species now or in the foreseeable future.

Factor C. Disease or Predation

    We do not have any information indicating that disease impacts 
Astragalus hamiltonii. We also do not have information on the effects 
of herbivory (eating) by livestock (see the Livestock Grazing section 
above), wildlife, or insects. However, we do not

[[Page 10172]]

believe herbivory from livestock is a concern due to the steepness of 
the terrain on which the plant is located and the time of year grazing 
occurs in A. hamiltonii habitat (see Factor A, Livestock Grazing). 
Based on the best available information, we do not believe A. 
hamiltonii is threatened by disease or predation now or for the 
foreseeable future.

Factor D. The Inadequacy of Existing Regulatory Mechanisms

    There are no laws protecting plants on private, State, or Tribal 
lands in Utah. A third of Astragalus hamiltonii individuals are found 
on BLM land. A. hamiltonii is listed as a bureau sensitive plant for 
the BLM. Limited policy-level protection by the BLM is afforded through 
the Special Status Species Management Policy Manual  6840 
which forms the basis for special status species management on BLM 
lands (BLM 2008e, entire).
    Despite the lack of regulatory mechanisms to protect Astragalus 
hamiltonii, we found that there are no threats to the species (Factors 
A, B, C, and E) that require regulatory mechanisms to protect the 
species. Therefore, we do not consider the inadequacy of regulatory 
mechanisms a threat to this species now or for the foreseeable future.

Factor E. Other Natural or Manmade Factors Affecting Its Continued 
Existence

    Natural and manmade factors affecting Astragalus hamiltonii 
include: (1) Small population size and (2) climate change and drought.
(1) Small Population Size
    We lack information on the population genetics of Astragalus 
hamiltonii, and as a probable outcrosser, this species could 
potentially be subject to the negative effects of small population 
size. As previously described (see Life History, above), plants that 
are obligate outcrossers cannot fertilize themselves and rely on other 
individual plants of differing genetic make-up to reproduce (Stebbins, 
1970, p. 310). Therefore, the fewer plants that are located at a site 
(i.e., small population size), the less chance exists for sufficient 
cross-fertilization.
    Small populations and species with limited distributions are 
vulnerable to relatively minor environmental disturbances (Given 1994, 
pp. 66-67). Small populations also are at an increased risk of 
extinction due to the potential for inbreeding depression, loss of 
genetic diversity, and lower sexual reproduction rates (Ellstrand and 
Elam 1993, entire; Wilcock and Neiland 2002, p. 275). Lower genetic 
diversity may, in turn, lead to even smaller populations by decreasing 
the species' ability to adapt, thereby increasing the probability of 
population extinction (Barrett and Kohn 1991, pp. 4, 28; Newman and 
Pilson 1997, p. 360).
    We do not believe small population size is a concern for Astragalus 
hamiltonii. A. hamiltonii grows robustly and in high densities with 
many flowers and fruits (Goodrich 2010b, entire; Goodrich 2010c, p. 
26). Although the species exists in a relatively small area (known 
distribution is 200 square miles (mi\2\) (518 square kilometers 
(km\2\)), it occurs across its range in a scattered--and potentially 
continuous--distribution. There are also large areas of suitable 
habitat that remain unsurveyed, so the species may be more widely 
distributed.
    Astragalus hamiltonii's scattered distribution may contribute to 
its overall viability and potential resilience (Goodrich 2010b, p. 89). 
For example, small-scale stochastic events, such as the erosion of a 
hillside during a flood event, would probably destroy only a small 
portion of the known individuals of A. hamiltonii. It is possible that 
a landscape-level event, such as a wildfire, could destroy most known 
A. hamiltonii individuals, but the sparseness of the vegetation and the 
lack of fine fuels in A. hamiltonii habitat makes this event unlikely 
(Wright and Bailey 1982, p. 1; Olmstead 2010, pers. comm.). The lack of 
other surface-disturbing threats (see Factor A) also leads us to 
believe that the species' current distribution and population size will 
remain intact.
    In the absence of information identifying threats to the species 
and linking those threats to the rarity of the species, we do not 
consider rarity alone to be a threat. A species that has always been 
rare, yet continues to survive, could be well equipped to continue to 
exist into the future. This may be particularly true for Astragalus 
hamiltonii, which is adapted to recolonize disturbed sites. Many 
naturally rare species have persisted for long periods within small 
geographic areas, and many naturally rare species exhibit traits that 
allow them to persist, despite their small population sizes. 
Consequently, the fact that a species is rare does not necessarily 
indicate that it may be in danger of extinction in the foreseeable 
future.
    Based on Astragalus hamiltonii's apparently robust reproductive 
effort, scattered distribution, and lack of other threats, we believe 
that small population size is not a threat to this species now or for 
the foreseeable future.
(2) Climate Change and Drought
    Climate change is likely to affect the long-term survival and 
distribution of native species, such as Astragalus hamiltonii, through 
changes in temperature and precipitation. Hot extremes, heat waves, and 
heavy precipitation will increase in frequency, with the Southwest 
experiencing the greatest temperature increase in the continental 
United States (Karl et al. 2009, pp. 28, 129). Approximately 20 to 30 
percent of plant and animal species are at increased risk of extinction 
if increases in global average temperature exceed 2.7 to 4.5 degrees 
Fahrenheit ([deg]F) (1.5 to 2.5 degrees Celsius ([deg]C)) 
(Intergovernmental Panel on Climate Change (IPCC) 2007, p. 48). In the 
southwestern United States, average temperatures increased 
approximately 1.5 [deg]F (0.8 [deg]C) compared to a 1960 to 1979 
baseline (Karl et al. 2009, p. 129). By the end of this century, 
temperatures are expected to warm a total of 4 to 10 [deg]F (2 to 5 
[deg]C) in the Southwest (Karl et al. 2009, p. 129).
    Annual mean precipitation levels are expected to decrease in 
western North America and especially the southwestern States by mid 
century (IPCC 2007, p. 8; Seager et al. 2007, p. 1181). Throughout 
Astragalus hamiltonii's range, precipitation is predicted to increase 
10 to 15 percent in the winter, decrease 5 to 15 percent in spring and 
summer, and remain unchanged in the fall under the highest emissions 
scenario (Karl et al. 2009, p. 29). The levels of aridity of recent 
drought conditions and perhaps those of the 1950s drought years will 
become the new climatology for the southwestern United States (Seager 
et al. 2007, p. 1181). Much of the Southwest remains in a 10-year 
drought, ``the most severe western drought of the last 110 years'' 
(Karl et al. 2009, p. 130). Although droughts occur more frequently in 
areas with minimal precipitation, even a slight reduction from normal 
precipitation may lead to severe reductions in plant production. 
Therefore, the smallest change in environmental factors, especially 
precipitation, plays a decisive role in plant survival in arid regions 
(Herbel et al. 1972, p. 1084).
    Atmospheric levels of carbon dioxide are expected to double before 
the end of the 21st century, which may increase the dominance of 
invasive grasses leading to increased fire frequency and severity 
across western North America (Brooks and Pyke 2002, p. 3; IPCC 2002, p. 
32; Walther et al. 2002, p. 391).

[[Page 10173]]

Elevated levels of carbon dioxide lead to increased invasive annual 
plant biomass, invasive seed production, and pest outbreaks (Smith et 
al. 2000, pp. 80-81; IPCC 2002, pp. 18, 32; Ziska et al. 2005, p. 1328) 
and will put additional stressors on rare plants already suffering from 
the effects of elevated temperatures and drought.
    No population trend data are available for Astragalus hamiltonii, 
but drought conditions led to a noticeable decline in survival, vigor, 
and reproductive output of other rare plants in the Southwest during 
the drought years of 2001 through 2004 (Anderton 2002, p. 1; Van Buren 
and Harper 2002, p. 3; Van Buren and Harper 2004, entire; Hughes 2005, 
entire; Clark and Clark 2007, p. 6; Roth 2008a, entire; Roth 2008b, pp. 
3-4).
    As discussed in the Life History section above, Astragalus 
hamiltonii seedling establishment is probably correlated with rainfall 
(Heil and Melton 1995a, p. 14); therefore, reduced precipitation may 
reduce seedling establishment. Additionally, the relatively localized 
distribution of A. hamiltonii may make this species more susceptible to 
landscape-level stochastic extinction events, such as regional drought. 
Despite these potential vulnerabilities, A. hamiltonii appears well-
adapted to a dry climate and can quickly colonize after disturbance. 
Plants growing in high-stress landscapes are adapted to stress, and 
drought-adapted species may experience lower mortality during severe 
droughts (Gitlin et al. 2006, pp. 1477, 1484).
    In summary, climate change is affecting and will affect temperature 
and precipitation events in the future. We expect that Astragalus 
hamiltonii, like other narrow endemics, may be negatively affected by 
climate change related drought. However, we believe that A. 
hamiltonii's adaptation to growing in high-stress environments renders 
this species less susceptible to negative effects from climate change. 
Although we believe climate change will impact plants in the future, 
the available information is too speculative to determine the 
likelihood of this potential threat to A. hamiltonii. Therefore, based 
on the best scientific and commercial information available, we 
conclude that climate change is not a threat to A. hamiltonii now or 
for the foreseeable future.
Summary of Factor E
    We assessed the potential risks of small population size, climate 
change, and drought to Astragalus hamiltonii. There is no evidence that 
the species' small population size is a threat to A. hamiltonii. 
Rather, small, scattered populations are likely an evolutionary 
adaptation of this species. Climate change and resulting drought may 
affect A. hamiltonii's growth and reproductive success. However, A. 
hamiltonii is adapted to a landscape where drought naturally occurs and 
is able to rapidly colonize after disturbance. In addition, as 
described in Factor A, there are no threats to the species that would 
result in significant loss or fragmentation of available habitat, and 
thus there are no cumulative effects to exacerbate the threat of 
climate change. We currently lack sufficient information that other 
natural or manmade factors rise to the level of a threat to A. 
hamiltonii now or for the foreseeable future.

Finding

    As required by the ESA, we conducted a review of the status of the 
species and considered the five factors in assessing whether Astragalus 
hamiltonii is endangered or threatened throughout all or a significant 
portion of its range. We examined the best scientific and commercial 
information available regarding the past, present, and future threats 
faced by A. hamiltonii. We reviewed the petition, information available 
in our files, and other available published and unpublished 
information, and we consulted with recognized A. hamiltonii experts and 
other Federal, State, and Tribal agencies.
    The primary factor potentially impacting Astragalus hamiltonii is 
future energy development (oil, gas, and tar sands). However, energy 
development is not likely to occur on a broad scale throughout this 
species' range in the foreseeable future. Furthermore, the best 
available information shows that A. hamiltonii can tolerate some 
habitat disturbances. Other factors affecting A. hamiltonii--including 
land conversion to agricultural use, grazing, recreation, nonnative 
invasive species, and small population size--are either limited in 
scope, or we do not have evidence that supports these factors adversely 
impacting the species as a whole. We have no evidence that 
overutilization, disease, and predation are affecting this species. 
Although climate change will likely impact plants in the future, we do 
not have enough information to determine that climate change will 
elicit a species-level response from A. hamiltonii. Finally, because 
none of these factors rises to the level of a threat, the inadequacy of 
regulatory mechanisms does not negatively affect A. hamiltonii.
    Based on our review of the best available scientific and commercial 
information pertaining to the five factors, we find that the factors 
analyzed above are not of sufficient imminence, intensity, or magnitude 
to indicate that Astragalus hamiltonii is in danger of extinction 
(endangered), or likely to become endangered within the foreseeable 
future (threatened), throughout its range. Therefore, we find that 
listing A. hamiltonii as a threatened or endangered species throughout 
its range is not warranted.

Significant Portion of the Range

    Having determined that Astragalus hamiltonii does not meet the 
definition of a threatened or endangered species, we must next consider 
whether there are any significant portions of the range where A. 
hamiltonii is in danger of extinction or is likely to become endangered 
in the foreseeable future.
    In determining whether a species is threatened or endangered in a 
significant portion of its range, we first identify any portions of the 
range of the species that warrant further consideration. The range of a 
species can theoretically be divided into portions an infinite number 
of ways. However, there is no purpose to analyzing portions of the 
range that are not reasonably likely to be significant and threatened 
or endangered. To identify only those portions that warrant further 
consideration, we determine whether there is substantial information 
indicating that: (1) The portions may be significant, and (2) the 
species may be in danger of extinction there or likely to become so 
within the foreseeable future. In practice, a key part of this analysis 
is whether the threats are geographically concentrated in some way. If 
the threats to the species are essentially uniform throughout its 
range, no portion is likely to warrant further consideration. Moreover, 
if any concentration of threats applies only to portions of the 
species' range that are not significant, such portions will not warrant 
further consideration.
    If we identify portions that warrant further consideration, we then 
determine whether the species is threatened or endangered in these 
portions of its range. Depending on the biology of the species, its 
range, and the threats it faces, the Service may address either the 
significance question or the status question first. Thus, if the 
Service considers significance first and determines that a portion of 
the range is not significant, the Service need not determine whether 
the species is threatened or endangered there. Likewise, if the Service 
considers status first and determines that the species is not 
threatened or endangered in a

[[Page 10174]]

portion of its range, the Service need not determine if that portion is 
significant. However, if the Service determines that both a portion of 
the range of a species is significant and the species is threatened or 
endangered there, the Service will specify that portion of the range as 
threatened or endangered under section 4(c)(1) of the ESA.
    We have no evidence that any particular population or portion of 
the range of Astragalus hamiltonii is critical to the species' 
survival. Although population area 2 appears to have a majority of the 
known Astragalus hamiltonii individuals, this area has received a 
majority of the search effort. A. hamiltonii may actually occur 
continuously across its known range, but range-wide surveys have not 
been done. The population areas delineated in this document were 
derived from existing data and information; however, information on the 
species' distribution and numbers may change with more survey effort. 
Additionally, potential threats to the species are essentially uniform 
throughout its range. Therefore, we do not find that A. hamiltonii is 
in danger of extinction now, nor is it likely to become endangered 
within the foreseeable future throughout all or a significant portion 
of its range. Therefore, listing A. hamiltonii as threatened or 
endangered under the ESA is not warranted at this time.
    We request that you submit any new information concerning the 
status of, or threats to, Astragalus hamiltonii to our Utah Ecological 
Services Field Office (see ADDRESSES section) whenever such information 
becomes available. New information will help us monitor A. hamiltonii 
and encourage its conservation. If an emergency situation develops for 
A. hamiltonii, or any other species, we will act to provide immediate 
protection.

Species Information--Penstemon flowersii

Taxonomy and Species Description
    Penstemon flowersii is an herbaceous plant in the figwort family 
(Scrophulariaceae) (Welsh et al. 2003, p. 624). This perennial plant 
can grow up to 14 in (36 cm) tall, with many branches that bloom dusty 
pink in May and June (Heil and Melton 1995b, pp. 6-7). It has dry, 
multi-part fruits less than 0.4 in (1 cm) long that split open when 
mature to release seeds (Neese and Welsh 1983, p. 429). P. flowersii 
has a poorly developed or absent basal rosette (a dense radiating 
cluster of leaves at the base of the plant) and smooth, thick leaves 
(Heil and Melton 1995b, pp. 6-7).
    Penstemon flowersii was first described in 1983 by Neese and Welsh, 
and is an accepted taxonomic entity (Welsh et al. 2003, p. 624). P. 
flowersii resembles other species in the genus and is closest 
vegetatively to P. carnosus (Heil and Melton 1995b, p. 8), but P. 
flowersii is distinguished by its smaller stature and dusty pink 
flowers (Neese and Welsh 1983, pp. 429-431). P. flowersii is closely 
related to P. immanifestus, a species that grows elsewhere in Nevada 
and Utah but has a more prominently bearded staminode (sterile male 
reproductive part found in the flower) (Heil and Melton 1995b, p. 8).
Distribution and Population Status
    Penstemon flowersii is found only in the Uinta Basin near 
Roosevelt, Utah. Its distribution straddles the Duchesne-Uintah County 
line (Figure 2). The species occurs across an area approximately 20 mi 
(32 km) by 4 mi (6.4 km) from Bridgeview to Randlett, Utah, in seven 
element occurrences (UNHP 2010b, entire) (see Distribution and 
Population Status section for Astragalus hamiltonii above for a 
complete definition of element occurrence). These seven element 
occurrences are not numbered consecutively because the UNHP combined 
previously disjunct element occurrences based on available information. 
As with A. hamiltonii, the element occurrences are recorded to the 
nearest quarter-quarter of the township, range, and section. This 
method of recording species locations gives the impression that element 
occurrences either overlap or join to form a continuous population. 
However, comprehensive surveys have not been done for all suitable 
habitats within an element occurrence, so we do not know if the 
population is continuous throughout the species' range.
    Penstemon flowersii was recently identified north of element 
occurrence 9 (Spencer 2010a, entire). We refer to this location as the 
``new site'' because it is not yet assigned to an element occurrence. 
At this time, we are unsure as to whether or not this new site will be 
designated as a new element occurrence or if it will be included in an 
existing element occurrence.
BILLING CODE 4310-55-P

[[Page 10175]]

[GRAPHIC] [TIFF OMITTED] TR23FE11.001

[FNP]BILLING CODE 4310-55-C
    Penstemon flowersii's distribution is patchy, although some sites 
can have moderately dense distribution with up to 10 plants in 1 yd\2\ 
(1 m\2\) (Heil and Melton 1995b, pp. 12-14). We do not know if the 
distribution of P. flowersii has changed over time because 
comprehensive surveys were not conducted for this species.
    Penstemon flowersii is found almost completely on private and 
Tribal lands (Table 2), with the exception of element occurrence 19, 
which is on property managed by the Utah Reclamation Mitigation and 
Conservation Commission for the U.S. Bureau of Reclamation (BOR) (UNHP 
2010b, entire).

[[Page 10176]]



                             Table 2--Estimated Number of Penstemon flowersii Plants
----------------------------------------------------------------------------------------------------------------
                                               Percent land ownership                                   Year of
           Element  occurrence           ---------------------------------     Number of penstemon        last
                                           Private     Tribal      BOR          flowersii plants         survey
----------------------------------------------------------------------------------------------------------------
1.......................................         75         25          0  2,000-13,000..............       2001
5.......................................         94          6          0  101-1,000.................       1995
6.......................................         78         22          0  No count..................       1982
8.......................................         71         29          0  61-71.....................       2004
9.......................................         91          9          0  51-100....................       2001
16......................................        100          0          0  4.........................       2001
19......................................         44         21         35  552.......................       2001
New site................................        100          0          0  29........................       2010
    Total...............................         79         19          2  2,798-14,756..............
----------------------------------------------------------------------------------------------------------------

    The total number of Penstemon flowersii individuals in Table 2 was 
derived from actual counts or estimates provided for each element 
occurrence. However, these counts do not include all known locations 
(e.g., private lands or BOR lands) for the species. The total number of 
P. flowersii individuals was previously estimated from 15,000 to 20,000 
on private lands alone, not including Tribal land (Heil and Melton 
1995b, p. 13; Franklin 2005, p. 131). We do not know how this estimate 
was derived.
    We cannot make a more accurate estimate for the total number of 
Penstemon flowersii because many sites on private and Tribal lands are 
inaccessible, and P. flowersii population numbers fluctuate widely from 
year to year (Heil and Melton 1995b, p. 16; Prevedel 2001 pers. comm. 
in Franklin 2005, p. 131). Therefore, we do not have accurate 
population counts or trend information for this species.
Habitat
    Penstemon flowersii is a narrow endemic that grows in Atriplex 
confertifolia (shadscale) communities on semibarren, gravelly clay 
slopes of the Uinta Formation (Heil and Melton 1995b, p. 9) at 
elevations ranging from 4,890 to 5,410 ft (1,490 to 1,650 m) 
(NatureServe 2009b, p. 2). It is found on both disturbed and 
undisturbed sites (Heil and Melton 1995b, p. 10).
Life History
    We know little of Penstemon flowersii's life history. Plant growth, 
seedling establishment, and juvenile mortality for this species are 
probably correlated with rainfall (Heil and Melton 1995b, p. 14). 
Reproduction and recruitment were noted at multiple sites across all 
element occurrences (UNHP 2010b, entire; Brunson 2010b, p. 1). One site 
had an estimated age structure of 4 percent seedlings and 96 percent 
mature adults, indicating that recruitment is occurring (UNHP 2010b, 
entire). Pollinators observed visiting P. flowersii include species of 
the order Hymenoptera: Anthophora affabilis, A. bomboides, and a 
species in the genus Osmia (Tepedino 2007, pers. comm. in Frates 2010, 
p. 32).

Summary of Information Pertaining to the Five Factors--Penstemon 
flowersii

    In making our 12-month finding on the petition, we considered and 
evaluated the best available scientific and commercial information 
pertaining to Penstemon flowersii in relation to the five factors 
provided in section 4(a)(1) of the ESA (see the full description of 
these five factors in the Summary of Information Pertaining to the Five 
Factors section for Astragalus hamiltonii above).

Factor A. The Present or Threatened Destruction, Modification, or 
Curtailment of Its Habitat or Range

    The following factors may affect the habitat or range of Penstemon 
flowersii: (1) Conversion to agricultural use/livestock grazing, (2) 
recreational activities, (3) oil and gas exploration and development, 
(4) nonnative invasive species, and (5) rural residential development.
(1) Conversion to Agricultural Use/Livestock Grazing
    For Penstemon flowersii, we combined two factors, conversion to 
agricultural use and livestock grazing, into one discussion because 
both of these factors occur on private lands. Historically, conversion 
of natural lands to agricultural use likely impacted Penstemon 
flowersii populations (Heil and Melton 1995b, pp. 8, 16), resulting in 
lower population numbers and habitat fragmentation. We believe the 
species was historically distributed in the low-lying areas because 
those areas that were not converted to agricultural use still contain 
P. flowersii plants (Franklin 2005, p. 131).
    Most of the suitable land in Duchesne and Uintah Counties was 
converted to agricultural use by 1970 (NAIP 2009, p. 2; Hilton 2010, 
p.1). Major changes in the amount of agricultural land in these 
counties are not expected in the future (Hilton 2010, p. 2). Therefore, 
we would not expect future agricultural conversion in these areas at a 
level that would threaten the species as a whole.
    The upper benches on private land where Penstemon flowersii now 
grows appear as nonirrigated terrain in digital imagery (NAIP 2009, p. 
2), and thus these areas are not likely used for agriculture. It is 
possible that most of these nonirrigated lands are used for rangeland 
grazing. Heavy grazing was noted at one site (UNHP 2010b, entire), and, 
as previously described, livestock can graze and trample plants (BLM 
2008c, p. 485). However, anecdotal observations indicate that this 
plant is not a preferred browse species by grazing livestock (Holmgren 
2009 pers. comm. in Frates 2010, p. 35), and the species can tolerate 
some level of soil disturbances (see Habitat). P. flowersii was noted 
as thriving in pastures (Holmgren 2009 pers. comm. in Frates 2010, p. 
35), so it appears that livestock grazing does not negatively impact 
the species. In summary, we have no information suggesting that 
conversion of habitat to agricultural use or livestock grazing are 
threats to P. flowersii now or for the foreseeable future.
(2) Recreational Activities
    Recreational activities (e.g., mountain bikes and motorized bikes) 
and OHV use can impact Penstemon flowersii and its habitat. The OHV use 
was documented within three element occurrences of P. flowersii to 
varying degrees (UNHP 2010b, entire). Two of these sites were listed in 
marginal condition, although plant vigor and reproduction at these 
sites was good (UNHP 2010b, entire). Disturbance occurred at a third 
site in 1995, and a population decline for this site was attributed to 
OHV activity (Heil and Melton 1995b, p. 17). However, vigorous plants 
were observed at this site with

[[Page 10177]]

ample flower production (UNHP 2010b, entire; Brunson 2010b, p. 1). The 
OHV use was not documented for the five remaining element occurrences 
or in the new P. flowersii site, but this does not necessarily mean OHV 
use does not occur there. Additionally, no other recreational uses were 
documented at P. flowersii sites.
    In summary, OHV use may be negatively affecting individual plants 
at some sites, but this impact is localized and not rangewide. We 
identified OHV use in the species' habitat, but the plants are vigorous 
and retaining their ability to reproduce. Therefore, we believe that 
recreational activities are not threats to Penstemon flowersii now or 
for the foreseeable future.
(3) Oil and Gas Exploration and Development
    Oil and gas exploration and development can impact Penstemon 
flowersii plants and their habitat (BLM 2008c, pp. 448-449). Within all 
mapped element occurrences of P. flowersii, there are four plugged and 
abandoned wells. All existing wells were plugged prior to 1999. As 
mentioned previously, plugged and abandoned wells involve surface 
disturbance for roads and well pads when they are constructed and 
during operation, but when they are abandoned they are reclaimed and do 
not receive regular traffic or disturbance (see Astragalus hamiltonii, 
Factor A, Oil and Gas Exploration and Development). There are two new 
proposed well locations within the species' mapped element 
occurrences--one well location that has an approved permit to drill and 
one well location that is not yet approved. Approved permits allow for 
well drilling, which will have associated negative impacts to 
vegetation, and potentially P. flowersii, during construction and 
drilling operations. These impacts have historically been localized and 
small in scale. We expect these impacts to continue to be minimal, 
considering that oil and gas development has occurred only minimally in 
P. flowersii habitat.
    The lack of oil and gas development in Penstemon flowersii habitat 
is most likely because there is not enough product to be economically 
feasible with current technology (Doyle 2010, pers. comm.; Sparger 
2010, pers. comm.) rendering dense energy developments unlikely in this 
area (BLM 2008c, p. 486). Although oil and gas development could 
potentially expand throughout P. flowersii habitat, substantial 
development is not likely for the next 20 years (BLM 2008c, p. 486), 
nor is it likely to occur across the entire range of P. flowersii. 
Thus, oil and gas exploration and development is not a threat to P. 
flowersii now or in the foreseeable future.
(4) Nonnative Invasive Species
    We have limited information regarding the distribution of nonnative 
invasive species in Penstemon flowersii habitat. We know that invasive 
species, particularly Bromus tectorum, occur within P. flowersii 
habitat (Frates 2010, pp. 29-30). However, we do not have any 
information indicating that B. tectorum or other nonnative invasive 
species impact P. flowersii.
    Soil disturbances can increase invasive species (see Astragalus 
hamiltonii, Factor A, Nonnative Invasive Species) (Evans et al. 2001, 
p. 1308). As noted above, B. tectorum, a major invasive plant species 
in the West, invades areas in response to surface disturbances (Hobbs 
1989, pp. 389, 393, 395, 398; Rejmanek 1989, pp. 381-383; Hobbs and 
Huenneke 1992, pp. 324-325, 329, 330; Evans et al. 2001, p. 1308). 
Therefore, we assessed the potential for soil disturbances to increase 
nonnative invasive species in the foreseeable future in Penstemon 
flowersii habitat.
    Agricultural use, livestock grazing, and oil and gas exploration 
and development are the predominant activities that disturb soils 
across the range of Penstemon flowersii. We determined that these 
activities are not extensive enough to threaten P. flowersii now or in 
the foreseeable future (see Agricultural Use/Livestock Grazing and Oil 
and Gas Exploration and Development). Thus, we also do not expect that 
these activities will increase surface disturbance to the point where 
invasive species will become established and impact P. flowersii to a 
significant degree. At this time, we have no information suggesting 
that nonnative invasive species are a threat to P. flowersii now or for 
the foreseeable future.
(5) Rural Residential Development
    Conversion of land for rural residential development can result in 
the permanent loss and fragmentation of habitat for many species, 
including Penstemon flowersii. Impacts include, but are not limited to, 
crushed vegetation, compacted soils, introduced exotic plant species, 
reduced available habitat, and increased habitat fragmentation (Hansen 
et al. 2005, entire). For the purpose of this analysis, we define rural 
residential development as the expansion of rural towns and surrounding 
rural areas through low-density housing construction and related 
business and industrial development.
    Duchesne and Uintah Counties, where Penstemon flowersii is found, 
had the highest (3.6 percent) and fourth highest (1.8 percent) 
population growth rates in Utah from 2008 to 2009, respectively (Utah 
Population Estimates Committee 2009, p. 2). The average population 
increase across the state of Utah was 1.5 percent over the same 
timeframe (Utah Population Estimates Committee 2009, p. 4). Roosevelt 
is the largest municipality that occurs near known P. flowersii 
habitat, and two smaller municipalities, Ballard and Myton, are nearby. 
The U.S. Census Bureau estimates that the population of Roosevelt 
increased approximately 12 percent from 2000 to 2009, with Ballard and 
Myton increasing 34 and 17 percent, respectively (U.S. Census Bureau 
2010a, entire). Human population growth can destroy and fragment 
habitat as municipalities grow and incorporate more of what was once 
natural land.
    Over the next 50 years, Duchesne and Uintah Counties are projected 
to grow at a slower rate of 1.1 percent (Utah Governor's Office of 
Planning and Budget (Utah GOPB) 2008, entire). At this growth rate, 
Daggett, Duchesne, and Uintah Counties (which are grouped together by 
the Utah Population Estimates Committee) are expected to increase from 
a current total population of 49,707 to 80,319 by 2060 (Utah GOPB 2008, 
entire). The City of Roosevelt projects a population of 6,600 by 2030, 
but they anticipate the population could be higher (City of Roosevelt 
2010, p. 7). Much of the urban and rural development in the Uinta Basin 
is influenced by the boom and bust cycles of energy development, and 
another boom cycle could increase population growth over predictions.
    Although municipalities are growing and are projected to increase 
near Penstemon flowersii habitat, they are not likely to impact a 
substantial amount of the known habitat of this species. The southern 
edge of Roosevelt's municipal boundary is approximately 0.2 mi (0.3 km) 
north of the northern boundary of element occurrence 1 (see Figure 2). 
The city limits of Ballard and Myton are immediately adjacent to 
element occurrences 1 and 9, with Ballard city limits overlapping 
element occurrence 6. None of these municipalities overlap with known 
sites of P. flowersii. Roosevelt will likely expand into an area 
already defined as an annexation area (City of Roosevelt 2010, p. 42), 
and this area is approximately 2 mi (3.2 km) north of element 
occurrence 9 and the

[[Page 10178]]

new site of P. flowersii on private land. Roosevelt and Ballard city 
limits are constrained by geography and Tribal boundaries, and neither 
are likely to expand substantially southward toward known P. flowersii 
sites (Eschler 2010, pers. comm.; Hyde 2010, pers. comm.).
    In summary, rural residential development is occurring now and is 
likely to increase in the future, but most of this development would 
occur outside of Penstemon flowersii known sites. Therefore, we do not 
believe rural residential development is a significant threat to the 
species now or in the foreseeable future.
Summary of Factor A
    Based on the best available information, we do not believe that 
conversion to agricultural use/livestock grazing, recreational 
activities, nonnative invasive species, oil and gas exploration and 
development, or rural residential development threaten Penstemon 
flowersii now or in the foreseeable future. Conversion to agricultural 
use most likely had an appreciable negative impact on P. flowersii 
historically, but we have no evidence that conversion to agricultural 
use continues today at a level that threatens the species. Likewise, 
livestock grazing is not widely noted across P. flowersii sites, and 
where it occurs it does not appear to negatively impact individuals. 
The OHV use, the only documented recreational activity in P. 
flowersii's habitat, is localized, and we do not have evidence that P. 
flowersii is considerably compromised or threatened by OHV use. We do 
not have information to support that nonnative invasive species are 
currently threatening P. flowersii or will be likely to do so in the 
foreseeable future. It is unlikely that current technologies and 
economic conditions will support substantial oil and gas development 
across P. flowersii habitat in the foreseeable future. Finally, rural 
residential development is unlikely to expand substantially into P. 
flowersii habitat. We find that the present or threatened destruction, 
modification, or curtailment of its habitat or range is not a threat to 
P. flowersii now or for the foreseeable future.

Factor B. Overutilization for Commercial, Recreational, Scientific, or 
Educational Purposes

    We are not aware of threats from overutilization or collection of 
Penstemon flowersii for commercial, recreational, scientific, or 
educational purposes, nor do we expect overutilization in the 
foreseeable future. P. duchesnensis, which is geographically near P. 
flowersii, is used horticulturally (Frates 2010, p. 75). However, P. 
flowersii is more obscure, and we have no evidence that this species is 
sought out for horticultural purposes (Frates 2010, p. 75). Therefore, 
we do not consider overutilization a threat to P. flowersii now or in 
the foreseeable future.

Factor C. Disease or Predation

    Disease and herbivory by insects, wildlife, or livestock was 
documented for Penstemon flowersii on only one occasion: Caterpillars 
were feeding on P. flowersii plants near Midview Reservoir (Spencer 
2010b, pers. comm.). We do not know how widespread this herbivory was 
or if it had detrimental effects on P. flowersii; caterpillars 
naturally feed on many plant species. The UNHP data did not note 
disease or herbivory for the species (UNHP 2010b, entire). With no data 
indicating otherwise, we do not consider disease or predation to be a 
threat to P. flowersii now or in the foreseeable future.

Factor D. The Inadequacy of Existing Regulatory Mechanisms

    There are no Federal or State laws that protect Penstemon 
flowersii. P. flowersii is found mostly on non-Federal lands, where no 
known regulatory mechanisms exist. However, we found that there are no 
threats to the species that warrant additional regulatory mechanisms 
(see Factors A, B, C, and E). Therefore, we do not consider the 
inadequacy of existing regulatory mechanisms as a threat to this 
species now or in the foreseeable future.

Factor E. Other Natural or Manmade Factors Affecting Its Continued 
Existence

    Natural and manmade threats to Penstemon flowersii's survival 
include: (1) Small population size and (2) climate change and drought.
(1) Small Population Size
    Penstemon flowersii grows across an area of 80 mi\2\ (207 km\2\). 
P. flowersii individuals occur in well-defined populations that are 
geographically isolated from one another. Thus, this species may be 
prone to the negative effects of small population size, in part because 
historical fragmentation of habitat (e.g., agricultural use) may have 
resulted in small populations with limited gene flow. P. flowersii also 
appears to have episodic growth patterns with large fluctuations in 
numbers from year to year (Franklin 2005, p. 131; 2010, p. 79). This 
fluctuation and patchy distribution may make P. flowersii more 
vulnerable to the impacts of small population size, limiting its 
ability to survive periods of low growth or recruitment.
    The species' biology, distribution, and even our information gaps 
indicate that small population sizes may not significantly impact 
Penstemon flowersii. For example, P. flowersii grows vigorously and in 
moderate densities with evidence of good reproduction and recruitment 
(UNHP 2010b, entire; Brunson 2010b, p. 1). Although we still consider 
P. flowersii a narrow endemic, it occurs across a relatively large 
range. In addition, there are relatively large amounts of unsurveyed 
potential habitat between known sites that could result in an expanded 
species distribution and range.
    Finally, we have not identified other surface-disturbing threats to 
this species that would cumulatively increase the risk of small 
population size. As previously discussed under Factor E for Astragalus 
hamiltonii (above), with no threats linked to a species' rarity, we do 
not consider rarity alone to be a threat. A species that has always 
been rare, yet continues to survive, could be well equipped to continue 
to exist into the future. Many naturally rare species have persisted 
for long periods within small geographic areas, and many naturally rare 
species exhibit traits that allow them to persist despite their small 
population sizes. Consequently, the fact that a species is rare does 
not necessarily indicate that it may be in danger of extinction in the 
foreseeable future. Thus, we believe that small population size is not 
a threat to P. flowersii.
(2) Climate Change and Drought
    Potential impacts of climate change and drought to the geographic 
area are characterized in the Climate Change and Drought section under 
Factor E for Astragalus hamiltonii (above). Penstemon flowersii occurs 
within the same geographic vicinity as A. hamiltonii and, therefore, 
will be exposed to similar changes in climate and drought.
    No trend data are available for Penstemon flowersii that would 
elucidate the relationship between the species' stability and climate 
variables. We do not know what causes fluctuations in P. flowersii 
abundance, but if it is due to environmental factors like precipitation 
or temperature, climate change could negatively affect this species. 
However, because of the lack of available data, any predictions are 
speculative.

[[Page 10179]]

    We expect that Penstemon flowersii, like other narrow endemics, may 
be negatively affected by climate change and drought. However, despite 
climate changes that have occurred over the past 30 years, we have no 
evidence that P. flowersii populations are declining, and we have no 
basis to predict how this species will respond in the future to climate 
change. Over the past 30 years, plant health remains normal to 
vigorous, and reproduction and recruitment continue to occur at some P. 
flowersii element occurrences (UNHP 2010b, entire). We have not 
identified other threats to this species, such as mining, that would 
cumulatively exacerbate the threat of climate change. Based upon the 
best available information, we do not believe that climate change is a 
threat now or is likely to become one in the foreseeable future.
Summary of Factor E
    We assessed the potential risks of small population size, climate 
change, and drought to Penstemon flowersii. There is no evidence that 
the species' small population size is a threat to P. flowersii. The 
species is adapted to a landscape where drought naturally occurs, and 
we have no information indicating that the species is threatened by 
climate change. In addition, as described in Factor A, there are no 
threats to the species that would result in significant loss or 
fragmentation of available habitat, and thus there are no cumulative 
effects to exacerbate the threat of climate change or small population 
sizes. Therefore, based on the best scientific and commercial 
information available at this time, we conclude that natural or manmade 
factors are not threats to P. flowersii now or for the foreseeable 
future.

Finding

    As required by the ESA, we conducted a review of the status of the 
species and considered the five factors in assessing whether Penstemon 
flowersii is endangered or threatened throughout all or a significant 
portion of its range. We examined the best scientific and commercial 
information available regarding the past, present, and future threats 
faced by P. flowersii. We reviewed the petition, information available 
in our files, other available published and unpublished information, 
and we consulted with recognized P. flowersii experts and other 
Federal, State, and Tribal agencies.
    The factor with potentially the most impact on Penstemon flowersii 
was historical agricultural development. Site visits show plants 
persist in pasture lands (Holmgren 2009 pers. comm. in Frates 2010, p. 
35; Brunson 2010b, p. 1), and we have little evidence that conversion 
to agricultural use is an ongoing threat to this species. Livestock do 
not appear to forage on P. flowersii, and the species occurs in grazing 
pastures. Rural residential development is another factor that could 
potentially destroy and fragment this species and its habitat in the 
future, but it is unlikely to occur at a high level across P. 
flowersii's known range. Other factors affecting P. flowersii--
including recreational activities, nonnative invasive species, oil and 
gas development, and small population size--are either limited in 
scope, or we do not have evidence that supports these factors adversely 
impacting the species as a whole. We have no evidence that 
overutilization, disease, and predation are affecting this species. 
Although climate change will likely impact the species, we do not have 
any information that indicates it threatens the continued existence of 
P. flowersii. Finally, because none of these factors rises to the level 
of a threat that would warrant additional regulatory mechanisms, the 
inadequacy of regulatory mechanisms does not negatively affect P. 
flowersii.
    Based on our review of the best available scientific and commercial 
information pertaining to the five factors, we find that the factors 
analyzed above are not of sufficient imminence, intensity, or magnitude 
to indicate that Penstemon flowersii is in danger of extinction 
(endangered), or likely to become endangered within the foreseeable 
future (threatened) throughout all or a significant portion of its 
range. Therefore, we find that listing P. flowersii as threatened or 
endangered species is not warranted throughout its range.

Significant Portion of the Range

    Having determined that Penstemon flowersii does not meet the 
definition of threatened or endangered species, we must next consider 
whether there are any significant portions of the range where P. 
flowersii is in danger of extinction or are likely to become endangered 
in the foreseeable future. See the Significant Portion of the Range 
section under Astragalus hamiltonii (above) for a summary of our 
interpretation of the meaning of ``in danger of extinction throughout 
all or a significant portion of its range.''
    We have no evidence that any particular population or portion of 
the range of Penstemon flowersii is critical to the species' survival. 
Because our understanding of the species' distribution is incomplete 
and population counts fluctuate widely, we cannot determine that any 
one element occurrence is more critical to the species' survival (i.e., 
has a significant portion of individuals) than another. Additionally, 
potential threats to the species appear to be uniform throughout P. 
flowersii's range. Therefore, we do not find that P. flowersii is in 
danger of extinction now, nor is it likely to become endangered within 
the foreseeable future throughout all or a significant portion of its 
range. Therefore, listing P. flowersii as threatened or endangered 
under the ESA is not warranted at this time.
    We request that you submit any new information concerning the 
status of, or threats to, Penstemon flowersii to our Utah Ecological 
Services Field Office (see ADDRESSES section) whenever such information 
becomes available. New information will help us monitor P. flowersii 
and encourage its conservation. If an emergency situation develops for 
P. flowersii, or any other species, we will act to provide immediate 
protection.

Species Information--Eriogonum soredium and Lepidium ostleri

    Eriogonum soredium and Lepidium ostleri occur in the same habitat 
and have the same distribution. Therefore, we discuss these species 
together for purposes of this finding.
Taxonomy and Species Description
    Eriogonum soredium
    Eriogonum soredium is a low mound-forming perennial plant in the 
buckwheat family (Polygonaceae) that is 0.8 to 1.6 in (2 to 4 cm) tall 
and 3.9 to 19.7 in (10 to 50 cm) across (Welsh et al. 2008, p. 588). 
The leaves are 0.08 to 0.2 in (2 to 5 mm) long, 0.03 to 0.08 in (0.7 to 
2 mm) wide, round to oval, and covered on both surfaces by short, 
white, wooly hairs (Welsh et al. 2008, p. 588). The numerous flowers 
are arranged in tight clusters resembling drumsticks. Individual 
flowers are white or partially pink and 0.08 to 0.12 in (2 to 3 mm) 
long (Welsh et al. 2008, p. 588). Flowering generally occurs from June 
to August. The seeds, which are 0.08 to 0.10 in (2 to 2.5 mm) long, 
mature from July through September (Welsh et al. 2008, p. 588).
    Eriogonum soredium was first described in 1981 by James Reveal 
based on a collection by Stan Welsh and Matt Chatterly (Reveal 1981, 
entire; Kass 1992a, p. 1). E. soredium has not undergone any taxonomic 
revisions since it was originally described. Therefore, we accept the 
current taxonomy as an indication that the

[[Page 10180]]

species constitutes a listable entity under the ESA.
Lepidium ostleri
    Lepidium ostleri is a long-lived perennial herb in the mustard 
family (Brassicaceae). It grows in dense cushion-like tufts up to 2 in 
(5 cm) tall (Welsh et al. 2008, p. 328). The grayish-green hairy leaves 
are 0.16 to 0.59 in (4 to 15 mm) long, generally linear, and entire or 
with lobed basal leaves (Welsh et al. 2008, p. 328). Flowering stalks 
are approximately 0.39 in (1 cm) long with 5 to 35 flowers that are 
white or have a purple tint (Welsh et al. 2008, p. 328). Flowering 
generally occurs from June to early July, followed by fruit set from 
July to August (Welsh et al. 2008, p. 328).
    Lepidium ostleri was first described in 1980 by Stan Welsh and 
Sherel Goodrich based on a collection by Stan Welsh and Matt Chatterly 
(Welsh and Goodrich 1980, entire; Kass 1992b, p. 1). L. ostleri has not 
undergone any taxonomic revisions since it was originally described. We 
are accepting the current taxonomy and consider L. ostleri a listable 
entity under the ESA.
Distribution and Population Status
    Eriogonum soredium and Lepidium ostleri are each known from four 
distinct, overlapping populations on private lands in the southern San 
Francisco Mountains in Beaver County, Utah--the Grampian Hill, Cupric 
Mine, Copper Gulch, and Indian Queen populations (Figure 3; Miller 
2010g, p. 6; Roth 2010a, pp. 1-2). We are not aware of any additional 
populations. Surveys were conducted on BLM lands adjacent to the known 
populations in 2010, and no plants or habitat were found (Miller 2010g, 
Appendix B and p. 6; Roth 2010a, pp. 1-3); these adjacent areas do not 
contain Ordovician Limestone, the substrate that supports both E. 
soredium and L. ostleri (see Habitat section below) (Miller 2010g, p. 
6). Similarly, no additional populations of either species were found 
during surveys of the San Francisco Mountains and surrounding ranges 
(including the Wah Wah Mountains, Crystal Peak, the Confusion Range, 
and the Mountain Home Range) (Kass 1992a, p. 5; Kass 1992b, p. 4; 
Evenden 1998, p. 5; Robinson 2004, p. 16; Miller 2010c, entire; Roth 
2010a, pp. 2-3).
    There were reports of two populations of E. soredium in the Wah Wah 
Mountains; however, we do not believe these reports are accurate--one 
report appears to have incorrect location information (Kass 1992a, p. 
5; Franklin 2005, p. 85) and the other report appears to be a species 
misidentification (Robinson 2004, p. 16; Roth 2010a, p. 3). Therefore, 
reports of these two populations are thought to be erroneous and are 
not discussed further in this finding.
BILLING CODE 4310-55-P

[[Page 10181]]

[GRAPHIC] [TIFF OMITTED] TR23FE11.002

BILLING CODE 4310-55-C
    Eriogonum soredium and Lepidium ostleri are distributed across a 
total range of less than 5 mi \2\ (13 km \2\). Previous estimates of 
the species' total occupied habitat ranged from 170 acres (ac) (69 
hectares (ha)) (Evenden 1998, Appendix C) to 400 ac (160 ha) (Kass 
1992a, pp. 7-8; 1992b, p. 7). However, we now have more accurate global 
positioning system information that shows the two species' total 
occupied habitat is approximately 52 ac (21 ha)

[[Page 10182]]

(based on Miller 2010g, Appendix B). For both species, each of the four 
known populations are estimated to occupy habitat ranging between 5 ac 
(2 ha) and 29 ac (12 ha), with localized high densities of plants 
(Evenden 1989, Appendix C; Miller 2010g, Appendix B).
    All known Eriogonum soredium and Lepidium ostleri populations are 
located on private lands (Miller 2010g, p. 6; Roth 2010a, pp. 1-2). 
Their occurrence on these private lands hinders our ability to collect 
accurate long-term population count or trend information because of 
access limitations. The populations were visited sporadically over the 
last couple of decades; however, we have no information on sampling 
methods used by individual surveyors. Common field techniques used to 
estimate population size tend to be highly subjective in the absence of 
actual population counts. Population estimates also may be skewed by 
how the species grow. Both species grow in low, mound-forming clusters, 
making it difficult to distinguish individual plants--some observers 
may assume each cluster is one plant and other observers might apply a 
multiplier to each cluster to count them as multiple plants; therefore, 
using either of these methods would greatly skew the resulting 
population estimate. We believe these biases help explain the seemingly 
large fluctuations in numbers of plants observed during different 
surveys (see below); E. soredium and L. ostleri are robust, long-lived 
perennial plants that are unlikely to exhibit such extreme population 
fluctuations (Garcia et al. 2008, pp. 260-261).
    Accordingly, the available population estimates are highly variable 
and probably not accurate. For Eriogonum soredium, available population 
estimates range from a low of 10 to 100 plants in 2004 to a high of 
76,000 to 81,000 individuals in 2010 (Kass 1992a, p. 8; Evenden 1998, 
Appendix C; Robinson 2004, pp. 11-15; Miller 2010a, pers. comm.; Miller 
2010b, pers. comm.; Miller 2010c, pp. 2-5; Roth 2010a, p. 4). For 
Lepidium ostleri, available population estimates range from a total of 
700 individuals (Kass 1992b, p. 8) to approximately 17,000 individuals 
in the 1990s (Evenden 1998, Appendix C). Currently, the total number of 
L. ostleri plants is estimated at approximately 43,000 (Miller 2010a, 
pers. comm.; Miller 2010c, pp. 2-5; Roth 2010a, p. 4). However, due to 
the aforementioned survey inaccuracies, we are not able to determine 
accurate population estimates or trends for either species. In 2010, 
both species were documented at all four known populations (Miller 
2010g, entire).
    We lack demographic information, which is measured by studying the 
size, distribution, composition, and changes within a specified 
population over time.
Habitat
    Eriogonum soredium and Lepidium ostleri are narrow endemics 
restricted to soils derived from Ordovician limestone outcrops (Evenden 
1998, p. 5). There are approximately 845 ac (342 ha) of Ordovician 
limestone outcrops in the San Francisco Mountains (Miller 2010g, 
Appendix F). In addition, there are 719 ac (291 ha) of Cambrian 
dolomite substrates in the San Francisco Mountains; there is the 
potential for small ``islands'' of Ordovician limestone outcrops to 
occur within these substrates (Miller 2010g, Appendix F, p. 7).
    Ordovician limestone is rare within a 50-mi (80-km) radius of the 
San Francisco Mountains (Miller 2010g, Appendix F). Cambrian dolomite 
substrates are present in the Wah Wah Mountains to the west of the San 
Francisco Mountains (Miller 2010g, Appendix F). However, as previously 
described (see Distribution and Population Status), there is no 
indication that additional populations of either species occur in these 
areas.
    We do not know if there are other limiting factors associated with 
the limestone formations that restrict the habitat use and distribution 
of these species--these species occupy only a fraction of the available 
habitat and are known to occur on only 52 ac (21 ha), or just 6 
percent, of the available Ordovician limestone outcrops.
    Eriogonum soredium and Lepidium ostleri are associated with pinion-
juniper and sagebrush communities between 6,200 and 7,228 ft (1,890 and 
2,203 m) in elevation. They are typically found on sparsely vegetated 
exposed slopes with Ephedra sp. (Mormon tea), Gutierrezia sarothrae 
(snakeweed), Cercocarpus intricatus (dwarf mountain-mahogany), and 
Petradoria pumila (rock goldenrod). Associated rare species include 
Trifolium friscanum.
Life History
    We do not have a clear understanding of the reproductive biology or 
life history of Eriogonum soredium, but recruitment appears to be low 
or perhaps episodic (Kass 1992a, p. 7; Roth 2010a, p. 1). Juvenile 
plants and seedlings have been observed in only two of the four 
populations (Miller 2010g, p. 4). In 2010, dead or partially dead 
plants were found throughout all populations, but we have no 
information on the cause of death or the approximate number of dead 
plants (Miller 2010g, p. 4).
    No information is available on the life history of Lepidium 
ostleri.

Summary of Information Pertaining to the Five Factors--Eriogonum 
soredium and Lepidium ostleri

    In making our 12-month finding on the petition, we considered and 
evaluated the best available scientific and commercial information 
pertaining to Eriogonum soredium and Lepidium ostleri in relation to 
the five factors provided in section 4(a)(1) of the ESA (see the full 
description of these five factors in the Summary of Information 
Pertaining to the Five Factors--Astragalus hamiltonii, above). E. 
soredium and L. ostleri co-occur in the same habitat and, therefore, 
are addressed together in the Five Factor Analysis below.

Factor A. The Present or Threatened Destruction, Modification, or 
Curtailment of Their Habitat or Range

    The following factors may affect the habitat or range of Eriogonum 
soredium and Lepidium ostleri: (1) Livestock grazing, (2) recreational 
activities, (3) mining, and (4) nonnative invasive species.
(1) Livestock Grazing
    Potential impacts of livestock grazing to plants are discussed 
above in the Livestock Grazing section under Factor A for Astragalus 
hamiltonii. As previously stated, all populations of Eriogonum soredium 
and Lepidium ostleri occur on private lands.
    We have no information on livestock grazing management on private 
lands, but adjacent BLM lands belong to active grazing allotments 
(Galbraith 2010, pers. comm.). Adjacent private lands are subject to 
the same grazing practices as the allotted BLM land if they are not 
fenced (Galbraith 2010, pers. comm.). Private lands in the San 
Francisco Mountains are only partially fenced; hence, livestock may 
have access to areas where E. soredium and L. ostleri occur. However, 
impacts to E. soredium or L. ostleri from livestock grazing have not 
been documented (Kass 1992a and 1992b, entire; Evenden 1998, entire; 
Miller 2010g, p. 5; Roth 2010a, p. 1).
    Based on our review of the available information, there is no 
indication that grazing impacts the species now or will impact the 
species in the foreseeable future at a level that threatens E. soredium 
or L. ostleri.
(2) Recreational Activities
    Potential impacts of recreational activities to plants are 
discussed above

[[Page 10183]]

in the Recreational Activities section under Factor A for Astragalus 
hamiltonii. There are no known impacts of OHV use in Eriogonum soredium 
and Lepidium ostleri occupied habitats (Miller 2010f, pers. comm.; Roth 
2010a, pp. 1-2). Access to the majority of the occupied habitat, which 
occurs on private lands, is posted as closed to all vehicles, including 
OHVs (Miller 2010g, p. 5). The OHV use does not appear to impact 
adjacent BLM lands in the San Francisco Mountains (Pontarolo 2009, 
pers. comm.). Therefore, we have no information indicating that 
recreational activities threaten E. soredium and L. ostleri now nor do 
we anticipate these activities will become a threat in the foreseeable 
future.
(3) Mining
    Mining activities occurred historically throughout the range of 
Eriogonum soredium and Lepidium ostleri and continue to impact these 
species. Mining activities can impact E. soredium and L. ostleri by 
removing habitat substrate, increasing erosion potential, fragmenting 
habitat through access road construction, degrading suitable habitat, 
and increasing invasive plant species (Brock and Green 2003, p. 15; BLM 
2008c, pp. 448-449). Impacts to E. soredium and L. ostleri individuals 
include crushing and removing plants, reducing plant vigor, and 
reducing reproductive potential through increased dust deposits, 
reduced seedbank quantity and quality, and decreased pollinator 
availability and habitat (Brock and Green 2003, p. 15; BLM 2008c, pp. 
448-449).
    The San Francisco Mountains have an extensive history of precious 
metal mining activity (Evenden 1998, p. 3). All four of the known 
populations and much of the species' potential habitat were impacted by 
precious metal mining activities in the past, as evidenced by a high 
density of mine shafts, tailings, and old mining roads throughout the 
habitat of Eriogonum soredium and Lepidium ostleri (Table 3; Kass 
1992a, p. 10; Evenden 1998, p. 3; Roth 2010a, p. 2).
    The eastern part of the Grampian Hill population surrounds old mine 
shafts associated with the King David Mine, which is part of the 
historical Horn Silver Mine. The Horn Silver Mine was one of the 
largest silver mines in the country until it collapsed in 1885 (Murphy 
1996, p. 1; Evenden 1998, p. 3). The Cupric Mine population is located 
immediately above a mine shaft associated with the Cupric Mine, a 
historical copper mine. Old mine shafts are located within 0.3 mi (0.5 
km) of the Copper Gulch population; these mine shafts are associated 
with the Cactus Mine, also a historical copper mine. Two mine shafts 
are located within the Indian Queen population and three additional 
mine shafts are located immediately adjacent to this population. These 
mine shafts also are part of the historical Cactus Mine.

              Table 3--Mining Activities in the Habitat of Eriogonum Soredium and Lepidium Ostleri
----------------------------------------------------------------------------------------------------------------
                                                                   Mining activity
            Population             -----------------------------------------------------------------------------
                                            Historical                 Current                   Future
----------------------------------------------------------------------------------------------------------------
Grampian Hill.....................  silver, lead, copper, zinc  None.................  silver, lead, copper,
                                     (Horn Silver Mine).                                zinc, landscape gravel
                                                                                        quarrying.
Cupric Mine.......................  silver, lead, copper,       gravel quarrying.....  silver, lead, copper,
                                     zinc, gravel quarrying                             zinc, landscape gravel
                                     (Cupric Mine).                                     quarrying.
Copper Gulch......................  silver, lead, copper,       gravel quarrying.....  silver, lead, copper,
                                     zinc, gravel quarrying                             zinc, landscape gravel
                                     (Cactus Mine).                                     quarrying.
Indian Queen......................  silver, lead, copper,       gravel quarrying.....  silver, lead, copper,
                                     zinc, gravel quarrying                             landscape gravel
                                     (Cactus Mine).                                     quarrying.
----------------------------------------------------------------------------------------------------------------

    Large-scale precious metal mining ceased decades ago. However, all 
precious metal mining claims in the southern San Francisco Mountains 
are patented (a claim for which the Federal Government has passed its 
title to the claimant, making it private land) and continued occasional 
explorations for silver, zinc, and copper deposits are reported for the 
area (Bon and Gloyn 1998, p. 12; Franconia Minerals Corporation 2002, 
p. 1; Rupke 2010, pers. comm.). In fact, in 1998 this area was one of 
the most active precious metal exploration areas in the State (Bon and 
Gloyn 1998, pp. 11-12). In addition, exploration activities were 
reinitiated at the Horn Silver Mine in 2002, confirming that extensive 
amounts of sphalerite (the major ore of zinc) remain in the mine 
(Franconia Minerals Corporation 2002, p. 1).
    We expect the demand for silver and copper to increase in the 
future (Crigger 2010, pp. 1-2; Murdoch 2010, pp. 1-2). The price for 
silver nearly tripled over the last decade (Stoker 2010, p. 2). The 
market for silver is expected to grow in the future due to its high 
demand for industrial uses in solar panel construction, wood 
preservatives, and medical supplies (Ash 2010, p. 1). Since 2009, the 
value of copper increased more than 140 percent (Crigger 2010, pp. 1-2; 
Murdoch 2010, pp. 1-2). The market for copper, one of the world's most 
widely used industrial metals, is expected to increase in the future 
due to demand for electrical wiring, plumbing, and car fabrication 
(Crigger 2010, pp. 1-2; Murdoch 2010, pp. 1-2). In Utah, precious 
metals accounted for approximately 14 percent of the total value of 
minerals produced in 2009 (up from 8 percent in 2008) (Utah GOPB 2010, 
pp. 195-196). Utah's precious metal gross production value increased 
$221 million (57 percent) compared to 2008, due to increased production 
of both gold and silver (Utah GOPB 2010, p. 196). Because the San 
Francisco Mountains area was one of the most productive areas during 
the last large-scale precious metal mining efforts, it is reasonable to 
assume that it will become important again, particularly given the 
ongoing exploration activities at the mines.
    As previously described, Eriogonum soredium and Lepidium ostleri 
are endemic to soils derived from Ordovician limestone. In addition to 
precious metals, this formation is mined for crushed limestone. The 
limestone is removed from quarry sites and sold for marble landscaping 
gravel.
    Marble landscaping gravel quarries in Eriogonum soredium and 
Lepidium ostleri's range are open-pit mines that result in the removal 
of the habitat substrate for these species. Four active limestone 
quarry sites occur within a couple hundred feet of three of the 
species' populations--Cupric Mine, Copper Gulch, and Indian Queen 
populations (Table 3).
    A limestone quarry is considered active from the time quarrying 
begins until the site is reclaimed. Generally, gravel pits are 
maintained below 5 ac (2 ha) of surface disturbance to avoid

[[Page 10184]]

large mine status, which requires permitting (Munson 2010, pers. 
comm.). Hence, an area may contain many quarries at or below the 5-ac 
(2-ha) threshold, all of which may be considered active (Munson 2010, 
pers. comm.). A mine also may stay below 5 ac (2 ha) as long as 
previously disturbed areas at the quarry site are reclaimed prior to 
expanding quarrying operations (Munson 2010, pers. comm.). The Cupric 
Mine, Copper Gulch, and Indian Queen populations of Eriogonum soredium 
and Lepidium ostleri all have small individual gravel pits--resulting 
in a lack of environmental analyses and potential mitigation 
opportunities (see Factor D, Inadequacy of Existing Regulatory 
Mechanisms).
    As stated in the Distribution and Population Status section above, 
Eriogonum soredium and Lepidium ostleri occur in the same overlapping 
locations, each occupying a total of 52 ac (21 ha) in four populations. 
We estimate the quarries at the three population sites (Cupric Mine, 
Copper Gulch, and Indian Queen) historically resulted in the loss of 26 
ac (11 ha) of suitable habitat adjacent to currently known plant 
locations (Table 4; Darnall et al. 2010, entire). Based on habitat 
similarities and proximity, it is likely that the plant occupied the 
entire 26 ac (11 ha) that are now being quarried. There are 23 ac (9 
ha) of remaining occupied habitat in the three populations (Table 4; 
Darnall et al. 2010, entire), but these areas are at risk of being 
impacted by the gravel pits. The only population not impacted by gravel 
pits--the Grampian Hill population--is 29 ac (12 ha) in size. Even so, 
the Grampian Hill population is only 1 mi (1.6 km) away from the 
nearest gravel pit and, as previously discussed, it is impacted by 
precious metal mining.

  Table 4--Areas of Surface Disturbance Associated With Gravel Mining in the Vicinity of Eriogonum Soredium and
                                          Lepidium Ostleri populations
----------------------------------------------------------------------------------------------------------------
              Population                      Occupied area               Adjacent surface  disturbance
---------------------------------------------------------------------------------------------------------------
Indian Queen..........................  9 ac (3.6 ha)............  14 ac (5.7 ha).
Copper Gulch..........................  5 ac (2.0 ha)............  5 ac (2.0 ha).
Cupric Mine...........................  9 ac (3.6 ha)............  7 ac (2.8 ha).
                                       -------------------------------------------------------------------------
    Total.............................  23 ac (9.2 ha)...........  26 ac (10.5 ha)............................
----------------------------------------------------------------------------------------------------------------

    Quarrying is occurring in the immediate vicinity of the Cupric Mine 
population (Evenden 1998, p. 5; Robinson 2004, p. 8; Frates 2006, pers. 
comm.; Roth 2010a, p. 2; Miller 2010e, pers. comm.; Munson 2010, pers. 
comm.); we anticipate this mining activity will continue to impact this 
population in the near future (Roth 2010a, p. 2). The estimated area of 
occupied habitat of the Cupric Mine population in the vicinity of this 
gravel pit is 9 ac (4 ha) (Table 4; Darnall et al. 2010, entire), while 
gravel mining has resulted in surface disturbance of approximately 7 ac 
(3 ha) (Table 4; Darnall et al. 2010, entire). No quarrying activity 
was observed in the vicinity of the Copper Gulch and Indian Queen 
populations in 2010; however, the gravel pits are still considered 
active and thus additional gravel mining could occur at any time. For 
both of these populations (Copper Gulch and Indian Queen), adjacent 
surface disturbance is equal to or greater than the remaining occupied 
habitat (Table 4; Darnall et al. 2010, entire).
    It is important to note that all of the active quarries are near or 
above the 5-ac (2-ha) regulatory limit. Thus, we anticipate that the 
operators will file for large mine permits, partially restore the 
disturbed areas to be below the 5-ac (2-ha) limit, or will begin new 
gravel pits (Munson 2010, pers. comm.). Under any of these scenarios, 
it is likely that occupied habitats of Eriogonum soredium and Lepidium 
ostleri will be impacted, particularly given the ongoing need for 
limestone gravel in nearby communities, as described below.
    Between 1995 and 2001, the production of building and landscaping 
stones in Utah jumped nearly 700 percent (Stark 2008, p. 1). 
Construction sand, gravel, and crushed stone production rank as the 
second most valuable commodity produced among industrial minerals in 
Utah (Bon and Krahulec 2009, p. 5). The use of landscape gravel will 
likely continue to increase in nearby Washington County, which is one 
of the fastest growing counties in the United States and Utah (U.S. 
Census Bureau 2010b, entire; Utah GOPB 2010, p. 48). The Washington 
County population has doubled every 10 years since 1970. In 2009, there 
were 145,466 people estimated to live in Washington County (Utah GOPB 
2010, p. 49). Over 700,000 people are expected to live in Washington 
County by 2050 (Utah GOPB 2008, entire). Based on the projected 
population growth for Washington County, we believe that the regional 
demand for landscape gravel will continue to increase in southwestern 
Utah in the foreseeable future.
    Much of the rock quarried in Utah does not travel far because of 
the associated high cost of transport (Stark 2008, p. 1). The quarries 
of the southern San Francisco Mountains are the closest quarries 
providing crushed limestone for southwestern Utah, including Washington 
County (Mine Safety and Health Administration 2010, p. 1). In addition 
to regional distribution, crushed limestone quarried from the vicinity 
of the Copper Gulch, Indian Queen, and Cupric Mine populations is 
transported to a distribution center for the Home Depot in the nearby 
town of Milford, where it is packaged and shipped nationwide (Munson 
2010, pers. comm.).
    To summarize, mining throughout Eriogonum soredium and Lepidium 
ostleri's range reduced available habitat and impacted the species' 
populations in the past (Table 3; Table 4). All four populations of 
Eriogonum soredium and Lepidium ostleri co-occur with precious metal 
mining activities. For both species, three of the four populations--the 
Cupric Mine, Copper Gulch, and Indian Queen populations--co-occur with 
active gravel mining pits.
    Available information suggests that all populations are likely to 
be impacted by precious metal and gravel mining in the foreseeable 
future based on mineral availability and market projections. Therefore, 
we have determined that mining is a threat to E. soredium and L. 
ostleri now and in the foreseeable future.
(4) Nonnative Invasive Species
    Potential impacts of nonnative invasive species to native plants 
and their habitat are discussed above in the Nonnative Invasive Species 
section under Factor A for Astragalus hamiltonii. Bromus tectorum is

[[Page 10185]]

considered the most ubiquitous invasive species in the Intermountain 
West due to its ability to rapidly invade native dryland ecosystems and 
outcompete native species (Mack 1981, p. 145; Mack and Pyke, 1983, p. 
88; Thill et al. 1984, p. 10).
    Bromus tectorum is a dominant species on the lower slopes of the 
Grampian Hill population and is present in all populations of Eriogonum 
soredium and Lepidium ostleri (Miller 2010g, p. 5; Roth 2010a, p. 1). 
Surface disturbances can increase the occurrence and densities of B. 
tectorum (see Nonnative Invasive Species section under Factor A for 
Astragalus hamiltonii). As previously described, increased mining 
activities and associated surface disturbances are expected to occur in 
the occupied habitat for E. soredium and L. ostleri, (see Mining, 
above), providing conditions allowing B. tectorum to expand into and 
increase density within E. soredium and L. ostleri habitat.
    Invasions of annual, nonnative species, such as Bromus tectorum, 
are well documented to contribute to increased fire frequencies (Brooks 
and Pyke 2002, p. 5; Grace et. al 2002, p. 43; Brooks et. al 2003, pp. 
4, 13, 15). The disturbance caused by increased fire frequencies 
creates favorable conditions for increased invasion by B. tectorum. The 
end result is a downward spiral where an increase in invasive species 
results in more fires, more fires create more disturbances, and more 
disturbances lead to increased invasive species densities. The risk of 
fire is expected to increase from 46 to 100 percent when the cover of 
B. tectorum increases from 12 to 45 percent or more (Link et al. 2006, 
p. 116). In the absence of exotic species, it is generally estimated 
that fire return intervals in xeric sagebrush communities range from 
100 to 350 years (Baker 2006, p. 181). In some areas of the Great Basin 
(Snake River Plain), fire return intervals due to B. tectorum invasion 
are now between 3 and 5 years (Whisenant 1990, p. 4). Most plant 
species occurring within a sagebrush ecosystem are not expected to be 
adapted to frequent fires, as evidenced in the lack of evolutionary 
adaptations found in other shrub-dominated fire adapted ecosystems like 
chaparral (Baker, in press, p. 17).
    In the absence of Bromus tectorum, Eriogonum soredium and Lepidium 
ostleri grow in sparsely vegetated communities unlikely to carry fires 
(see Habitat section). Thus, the species are unlikely to be adapted to 
survive fires. As described in the distribution section, the total 
range of these species are less than 5 mi\2\ (13 km\2\) and each of the 
four populations occupy relatively small areas ranging between 5 ac (2 
ha) and 29 ac (12 ha). A range fire could easily impact, or eliminate, 
one or all populations. Therefore, the potential expansion of invasive 
species and associated fire is a threat to the species, especially when 
considering the limited distribution of the species and the high 
potential of stochastic extinctions (as discussed in the Small 
Population Size section under Factor E below).
    In summary, nonnative invasive species and fire are threats to both 
species. Bromus tectorum occurs in all four Eriogonum soredium and 
Lepidium ostleri populations. Given the ubiquitous nature of B. 
tectorum in the Intermountain West and its ability to rapidly invade 
dryland ecosystems (Mack 1981, p. 145, Mack and Pyke, 1983, p. 88, 
Thill et al. 1984, p. 10), we expect it to increase in the future in 
response to surface disturbances from increased mining activities and 
global climate change (see the Climate Change and Drought section under 
Factor E for Astragalus hamiltonii). An increase in B. tectorum is 
expected to increase the frequency of fires in E. soredium and L. 
ostleri's habitat, and the species are unlikely to survive increased 
wildfires due to their small population sizes. Therefore, we determine 
that nonnative invasive species and associated wildfires constitute a 
threat to all populations of E. soredium and L. ostleri now and into 
the foreseeable future.
Summary of Factor A
    At this time, based on best available information, we do not 
believe that grazing and recreational activities significantly threaten 
Eriogonum soredium and Lepidium ostleri now or in the foreseeable 
future. However, we determine that mining and nonnative invasive 
species are threats to E. soredium and L. ostleri.
    Mining activities impacted Eriogonum soredium and Lepidium ostleri 
habitat in the past and continue to be a threat to the species and its 
habitat throughout its range. All of the populations and the majority 
of habitat are located on private lands with an extensive history and 
recent successful exploration activities for precious metal mining. 
Three of the four populations are located in the immediate vicinity of 
gravel mining. Gravel mining is expected to continue and expand in the 
near future (Munson 2010, pers. comm.). Considering the small acreages 
of occupied habitat immediately adjacent to existing gravel pits, 
continued mining may result in the loss of these populations in the 
foreseeable future. We anticipate an increase in the demand for 
precious metals and landscape rock based on the economic outlook for 
these commodities and the lack of alternative sources for crushed 
limestone in southwestern Utah which will result in increased impacts 
to E. soredium and L. ostleri and their habitat.
    Bromus tectorum is documented to occur in all four populations of 
Eriogonum soredium and Lepidium ostleri. The threat of fire caused by 
annual nonnative species invasions is exacerbated by mining activities 
and global climate change (see the Climate Change and Drought section 
under Factor E). The small population sizes and extremely limited 
distribution make this species especially vulnerable to stochastic 
extinction events, including localized mining activities and wildfires 
caused by increased invasions of nonnative species (see the Small 
Population Size section under Factor E, below).
    Therefore, we find that Eriogonum soredium and Lepidium ostleri are 
threatened by the present or threatened destruction, modification, or 
curtailment of the species' habitat or range, now and in the 
foreseeable future, based on impacts from mining activities and 
nonnative invasive species.

Factor B. Overutilization for Commercial, Recreational, Scientific, or 
Educational Purposes

    Eriogonum soredium and Lepidium ostleri are considered attractive 
rock garden plants. In particular, Eriogonum soredium is considered 
``one of the most fantastic of its genus'' by a major rock garden seed 
distributor (Alplains Seed Catalog 2010b, pp. 2 and 12). Seeds for both 
species are available commercially and they are harvested from wild 
populations (Alplains Seed Catalog 2010b, pp. 2 and 12).
    Eriogonum soredium and Lepidium ostleri plants are located on 
private lands, which may provide some protection from collectors, as 
access is restricted on these private lands. Despite the attractiveness 
of the two species to horticultural enthusiasts, we have no information 
indicating that collection in the wild is a threat to the species.
    In summary, overutilization for commercial purposes could be a 
concern to Eriogonum soredium and Lepidium ostleri due to their 
desirability to collectors; however, we do not have information that 
leads us to believe that overutilization for commercial purposes is a 
threat now or is likely to become one in the foreseeable future.

[[Page 10186]]

Factor C. Disease or Predation

    Disease and herbivory of the species are unknown. We do not have 
any information indicating that disease is impacting either Eriogonum 
soredium or Lepidium ostleri. We also do not have any information 
indicating herbivory is occurring from livestock (see the Livestock 
Grazing section under Factor A), wildlife, or insects (Kass 1992a, p. 
9; Evenden 1998, entire; Miller 2010a, entire; Miller 2010b, entire; 
Miller 2010c, entire; Roth 2010a, entire). Thus, we do not consider 
disease and predation to be threats to these species.

Factor D. The Inadequacy of Existing Regulatory Mechanisms

    There are no endangered species laws protecting plants on private, 
State, or Tribal lands in Utah. Eriogonum soredium and Lepidium ostleri 
are listed as bureau sensitive plants for the BLM. Should the species 
be located on BLM lands, limited policy-level protection by the BLM is 
afforded through the Special Status Species Management Policy Manual 
 6840, which forms the basis for special status species 
management on BLM lands (BLM 2008e, entire).
    Eriogonum soredium and Lepidium ostleri are predominantly 
threatened by mining related activities (see Factor A). Over 90 percent 
of the species' known potential habitat and all of the known 
populations are located on lands with private, patented mining claims 
(Kass 1992a, p. 9; Evenden 1998, p. 9; Roth 2010a, pp. 1-2). Mineral 
mining is subject to the Utah Mined Land Reclamation Act of 1975, which 
includes mineral mining on State and private lands, including lands 
with patented mining claims (Utah Code Title 40, Chapter 8). The ESA 
applies to all surface activities associated with the exploration, 
development, and extraction of mineral deposits.
    The Utah Mined Land Reclamation Act mandates the preparation of 
State environmental impact assessments for large mining operations, 
which are defined as mining operations which create more than 5 ac (2 
ha) of surface disturbance (UDOGM 2010b, p. 1). The existing gravel 
mining activities within the range of Eriogonum soredium and Lepidium 
ostleri (see Factor A, Mining) are approaching the 5-ac (2-ha) 
regulatory threshold. Thus, we anticipate that the operators will file 
for large mine permits, partially restore the disturbed areas to be 
below the 5-ac (2-ha) limit, or will begin new gravel pits (Munson 
2010, pers. comm.).
    State environmental impact assessments must address, at a minimum, 
the potential effects on State and federally listed species (Baker 
2010, pers. comm.). Eriogonum soredium and Lepidium ostleri are not 
State listed but are on the BLM sensitive species list. If UDOGM is 
made aware of these rare species being impacted by mining activities, 
they could consider minimizing and mitigating impacts; however, there 
is no requirement to address species that are not federally listed in 
the mine permitting process (Baker 2010, pers. comm.).
    In summary, the existing regulatory mechanisms are not adequate to 
protect Eriogonum soredium and Lepidium ostleri from becoming 
threatened or endangered by gravel mining on private lands. The active 
gravel pits are approaching the 5-ac (2-ha) threshold that would 
normally incur regulatory environmental impact assessments; however, no 
assessments are completed for these mines. Even if an environmental 
impact assessment is completed for any of the mines, the existing 
mining laws do not necessarily apply to BLM sensitive species: They 
recommend, and do not mandate, species protection or mitigation. Thus, 
we find that the inadequacy of existing mechanisms to regulate mining 
activities on private lands is a threat to all populations of E. 
soredium and L. ostleri now and in the foreseeable future.

Factor E. Other Natural or Manmade Factors Affecting Its Continued 
Existence

    Natural and manmade threats to Eriogonum soredium and Lepidium 
ostleri's survival include: (1) Small population size and (2) climate 
change and drought.
(1) Small Population Size
    General potential impacts of small population sizes to plants are 
discussed above in the Small Population Size section under Factor E for 
Astragalus hamiltonii.
    As previously described (see the Distribution and Population Status 
section), the entire ranges of both species are located in an area of 
less than 5 mi\2\ (13 km\2\). Within this range, each of the four 
individual populations' occupied habitat areas are very small, ranging 
from 5 ac (2 ha) to 29 ac (12 ha) (based on Miller 2010g, Appendix B).
    Eriogonum soredium and Lepidium ostleri can be dominant in small 
areas of occupied habitat, containing thousands of individuals. 
However, the small areas of occupation and the narrow overall range of 
the species make it highly susceptible to stochastic extinction events 
and the effects of inbreeding depression.
    Despite the overall lack of information on the population ecology 
of Eriogonum soredium and Lepidium ostleri, we know that small 
populations are at an increased risk of extinction due to the potential 
for inbreeding depression, loss of genetic diversity, and lower sexual 
reproduction rates (Ellstrand and Elam 1993, entire; Wilcock and 
Neiland 2002, p. 275). We do not have a clear understanding of the 
reproductive biology of E. soredium and L. ostleri, but recruitment 
appears to be low or episodic for E. soredium (Kass 1992a, p. 7; Roth 
2010a, p. 1). Low levels of recruitment in small populations may be due 
to inbreeding depression caused by the lack of genetic diversity and 
low levels of genetic exchange between populations (Ellstrand and Elam 
1993, entire; Wilcock and Neiland 2002, p. 275).
    Mining, or a single random event such as a wildfire (see Factor A), 
could extirpate an entire or substantial portion of a population given 
the small acreages of occupied habitat. Species with limited ranges and 
restricted habitat requirements also are more vulnerable to the effects 
of global climate change (see the Climate Change and Drought section 
below; IPCC 2002, p. 22; Jump and Penuelas 2005, p. 1016; Machinski et 
al. 2006, p. 226; Krause 2010, p. 79).
    Overall, we consider small population size an intrinsic 
vulnerability to Eriogonum soredium and Lepidium ostleri that may not 
rise to the level of a threat on its own. However, the small population 
sizes rise to the level of a threat because of the combined effects of 
small population sizes, limited distribution, and narrow overall range, 
compounded by the effects of global climate change (see below) and the 
potential for stochastic extinction events such as mining and invasive 
species (see Factor A). Therefore, we consider small localized 
population size, in combination with mining, invasive species, and 
climate change, to be a threat to both species now and in the 
foreseeable future.
(2) Climate Change and Drought
    Potential impacts of climate change and drought to the geographic 
area are characterized under Factor E for Astragalus hamiltonii. As 
discussed above, Eriogonum soredium and Lepidium ostleri have a limited 
distribution and populations are localized and small. In addition, 
these populations are restricted to very specific soil types. Global 
climate change exacerbates the risk of extinction for species that are 
already vulnerable due to low population numbers and restricted habitat 
requirements (see the

[[Page 10187]]

Climate Change and Drought section under Factor E for Astragalus 
hamiltonii).
    Predicted changes in climatic conditions include increases in 
temperature, decreases in rainfall, and increases in atmospheric carbon 
dioxide in the American Southwest (Walther et al. 2002, p. 389; IPCC 
2007, p. 48; Karl et al. 2009, p. 129). Although we have no information 
on how Eriogonum soredium and Lepidium ostleri will respond to effects 
related to climate change, persistent or prolonged drought conditions 
are likely to reduce the frequency and duration of flowering and 
germination events, lower the recruitment of individual plants, 
compromise the viability of populations, and impact pollinator 
availability (Tilman and El Haddi 1992, p. 263; Harrison 2001, p. 78). 
The smallest change in environmental factors, especially precipitation, 
plays a decisive role in plant survival in arid regions (Herbel et al. 
1972, p. 1084).
    Drought conditions led to a noticeable decline in survival, vigor, 
and reproductive output of other rare and endangered plants in the 
Southwest during the drought years of 2001 through 2004 (Anderton 2002, 
p. 1; Van Buren and Harper 2002, p. 3; Van Buren and Harper 2004, 
entire; Hughes 2005, entire; Clark and Clark 2007, p. 6; Roth 2008a, 
entire; Roth 2008b, pp. 3-4). Similar responses are anticipated to 
adversely affect the long-term persistence of E. soredium and L. 
ostleri.
    Climate change is expected to increase levels of carbon dioxide 
(Walther et al. 2002, p. 389; IPCC 2007, p. 48; Karl et al. 2009, p. 
129). Elevated levels of carbon dioxide lead to increased invasive 
annual plant biomass, invasive seed production, and pest outbreaks 
(Smith et al. 2000, pp. 80-81; IPCC 2002, pp. 18, 32; Ziska et al. 
2005, p. 1328) and will put additional stressors on rare plants already 
suffering from the effects of elevated temperatures and drought.
    The actual extent to which climate change itself will impact 
Eriogonum soredium and Lepidium ostleri is unclear, mostly because we 
do not have long-term demographic information that would allow us to 
predict the species' responses to changes in environmental conditions, 
including prolonged drought. Any predictions at this point on how 
climate change would affect these species would be speculative. 
However, as previously described, the species are threatened by mining 
activities (see Mining, Factor A) which will likely result in the loss 
of large numbers of individuals and maybe even entire populations. 
Increased surface disturbances associated with mining activities also 
will likely increase the extent and densities of nonnative invasive 
species and with it the frequencies of fires (see Nonnative Invasive 
Species section under Factor A). Given the cumulative effects of the 
potential population reduction and habitat loss (of already small 
populations) associated with mining, invasive species, and fire, we are 
concerned about the impacts of future climate change to Eriogonum 
soredium and Lepidium ostleri.
    In summary, we find it difficult to analyze the potential effects 
of global climate change on Eriogonum soredium and Lepidium ostleri in 
the absence of demographic trend data for the species which would allow 
us to analyze how they respond to climate change over time. However, 
because of the threats of mining, nonnative species, and small 
population size, the cumulative effects of climate change may be of 
concern for these species in the future. At this time, we believe that 
the state of knowledge concerning the localized effects of climate 
change is too speculative to determine whether climate change is a 
threat to these species in the foreseeable future. However, we will 
continue to assess the potential of climate change to threaten the 
species as better scientific information becomes available.
Summary of Factor E
    We assessed the potential risks of small population size, climate 
change, and drought to Eriogonum soredium and Lepidium ostleri 
populations. E. soredium and L. ostleri have a highly restricted 
distribution and exist in four populations scattered over an area that 
is less than 5 mi\2\ (13 km\2\). Individual populations occupy very 
small areas with large densities of plants. Even in the absence of 
information on genetic diversity, inbreeding depression, and 
reproductive effort, we believe a random stochastic event could impact 
a significant portion of a population. Small populations that are 
restricted by habitat requirements also are more vulnerable to the 
effects of climate change, such as prolonged droughts and increased 
fire frequencies.
    While naturally occurring droughts are not likely to impact the 
long-term persistence of the species, an increase in periodic prolonged 
droughts due to climate change could impact the species across their 
entire range in the future. Global climate change, particularly when 
assessed cumulatively with small population sizes and threats from 
mining activities, could increase the density of invasive annual 
plants, which are already present in the habitat of Eriogonum soredium 
and Lepidium ostleri (see Factor A). Increased nonnative species in the 
habitat of E. soredium and L. ostleri can increase fire frequency and 
severity. Because E. soredium and L. ostleri are not likely adapted to 
persist through fires, wildfires can have a significant impact on these 
small populations.
    Although small population size and climate change make the species 
intrinsically more vulnerable, we are uncertain whether they would rise 
to the level of threat by themselves. However, when combined with the 
threats listed under Factor A (mining and nonnative invasive species), 
small population size is likely to rise to the level of threat in the 
foreseeable future. At this time, we are uncertain of the degree to 
which climate change constitutes a threat to the species.

Finding

    As required by the ESA, we conducted a review of the status of the 
species and considered the five factors in assessing whether Eriogonum 
soredium and Lepidium ostleri are endangered or threatened throughout 
all or a significant portion of their range. We examined the best 
scientific and commercial information available regarding the past, 
present, and future threats faced by E. soredium and L. ostleri. We 
reviewed the petition, information available in our files, and other 
available published and unpublished information, and we consulted with 
E. soredium and L. ostleri experts and other Federal and State 
agencies.
    This status review identified threats to the species attributable 
to Factors A, D, and E. The primary threat to the species is habitat 
destruction from precious metal and gravel mining on private lands 
(Factor A). All populations are located in the vicinity of historical 
precious metal mining activities, at which ongoing exploration 
activities show the potential for continued mining activities in the 
foreseeable future. Three of the four populations are in the immediate 
vicinity of limestone quarries, all of which are considered active. We 
expect an increase in precious metal and limestone mining at these 
locations in the foreseeable future, with associated loss and 
fragmentation of Eriogonum soredium and Lepidium ostleri populations.
    Bromus tectorum occurs within all four Eriogonum soredium and 
Lepidium ostleri populations. It is a highly invasive nonnative species 
that spreads quickly in response to surface disturbances such as 
mining. As previously discussed, both species

[[Page 10188]]

occur in the immediate vicinity of precious metal and limestone mines--
mines inherently cause surface disturbances from excavation activities 
and the construction of roads and other infrastructure. Global climate 
change is expected to increase drought conditions in the Southwest and 
increase the spread of nonnative invasive species. The biggest concern 
associated with the increase in invasive species is the threat of 
increased wildfire (Factor A), particularly when considering the small 
population sizes and small occupied habitat area associated with these 
species.
    The magnitude of the biological threats posed by the species' small 
population sizes and limited ranges are not well understood due to the 
lack of information available on the ecology of Eriogonum soredium and 
Lepidium ostleri. Future studies may provide us with a more thorough 
understanding of threats posed by pollinator limitation, inbreeding 
depression, and the potential lack of genetic diversity over the 
species' range. However, the small areas of occupied habitat make the 
species highly vulnerable to habitat destruction through mining-related 
activities as well as random extinction events, including invasive 
species (and the inherent risk of increased fires) and the potential 
future effects of global climate change (Factor E).
    The existing regulatory mechanisms are not adequate to protect 
Eriogonum soredium and Lepidium ostleri from the primary threat of 
mining, particularly because both species occur entirely on private 
lands. The inadequacy of regulatory mechanisms (Factor D) on private 
land, combined with the economic and commercial value of the limestone 
and precious metals, poses a serious threat to the continued existence 
of E. soredium and L. ostleri. Ongoing mining in the habitat of E. 
soredium and L. ostleri has the potential to extirpate one of the four 
populations in the near future; all populations have the potential to 
be extirpated by mining-related activities in the foreseeable future 
(Factor A; Table 3).
    On the basis of the best scientific and commercial information 
available, we find that the petitioned action to list Eriogonum 
soredium and Lepidium ostleri as endangered or threatened is warranted. 
We will make a determination on the status of the species as endangered 
or threatened when we do a proposed listing determination. However, as 
explained in more detail below, an immediate proposal of a regulation 
implementing this action is precluded by higher priority listing 
actions, and progress is being made to add or remove qualified species 
from the Lists of Endangered and Threatened Wildlife and Plants.
    We reviewed the available information to determine if the existing 
and foreseeable threats render the species at risk of extinction now 
such that issuing an emergency regulation temporarily listing the 
species under section 4(b)(7) of the ESA is warranted. We determined 
that issuing an emergency regulation temporarily listing the species is 
not warranted at this time because there is no emergency posing a 
significant risk to the well-being of Eriogonum soredium or Lepidium 
ostleri. We do not believe that any of the potential threats are of 
such great immediacy and severity that would threaten all of the known 
populations with the imminent risk of extinction. However, if at any 
time we determine that issuing an emergency regulation temporarily 
listing Eriogonum soredium and Lepidium ostleri is warranted, we will 
initiate this action at that time.

Listing Priority Number

    The Service adopted guidelines on September 21, 1983 (48 FR 43098), 
to establish a system for utilizing available resources for the highest 
priority species when adding species to the Lists of Endangered or 
Threatened Wildlife and Plants or reclassifying species listed as 
threatened to endangered status. These guidelines, titled ``Endangered 
and Threatened Species Listing and Recovery Priority Guidelines,'' 
address the immediacy and magnitude of threats, as well as the level of 
taxonomic distinctiveness, by assigning priority in descending order to 
monotypic genera (genus with one species), full species, and subspecies 
(or equivalently, DPS of vertebrates). We assigned Eriogonum soredium 
and Lepidium ostleri each a Listing Priority Number (LPN) of 8, based 
on our finding that both species face threats of moderate magnitude 
that are imminent. These threats include the present or threatened 
destruction, modification or curtailment of their habitat, the 
inadequacy of existing regulatory mechanisms, and other manmade factors 
affecting their continued existence. These threats are ongoing and, in 
some cases (such as nonnative species), are considered irreversible, 
because, in the case of nonnative species invasions, large-scale 
invasions cannot be recovered to a native functioning ecosystem. Our 
rationale for assigning E. soredium and L. ostleri an LPN of 8 is 
outlined below.
    Under the Service's LPN Guidance, the magnitude of threat is the 
first criterion we look at when establishing a listing priority. The 
guidance indicates that species with the highest magnitude of threat 
are those species facing the greatest threats to their continued 
existence. These species receive the highest listing priority. We 
consider the threats that Eriogonum soredium and Lepidium ostleri face 
to be moderate in magnitude because the major threats (mining, 
nonnative species, small population size, climate change, and 
inadequacy of existing regulatory mechanisms), while serious and 
occurring rangewide, do not collectively rise to the level of high 
magnitude. For example, active mining is currently impacting only one 
of the four populations.
    The magnitude of Factor A is considered moderate, because, although 
we think that all populations have been impacted by mining in the past 
and three of the four populations occur in the immediate vicinity of 
gravel pits, mining activities are currently ongoing in one of these 
gravel pits. Ongoing mining in the habitat of E. soredium and L. 
ostleri is expected to increase the density of Bromus tectorum, thereby 
facilitating the spread of fire. B. tectorum is currently documented in 
all populations.
    We considered the magnitude of Factor D to be moderate. All 
populations are located on private lands with patented mining claims, 
where existing regulatory mechanisms are not adequate to protect 
Eriogonum soredium and Lepidium ostleri from the impacts of mining. All 
populations have the potential to be impacted by gravel and precious 
metal mining in the future; however, because only one population is 
currently impacted by gravel mining, we consider this threat to be 
moderate.
    We consider the magnitude of Factor E to be moderate, because 
although small population size and climate change make the species 
intrinsically more vulnerable, we are uncertain of whether they would 
rise to the level of threat by themselves. However, when collectively 
analyzed with the threats listed under Factor A, they may rise to the 
level of threat in the foreseeable future. Although we are uncertain 
about the direct impacts of global climate change on Eriogonum soredium 
and Lepidium ostleri, we expect the species to respond negatively to 
changed environmental conditions and drought, primarily from an 
increase in nonnative invasive species and wildfire (see Factor A). The 
threats of nonnative invasive species and wildfire could result in the 
extirpation of all populations, especially because the populations are 
small in size.

[[Page 10189]]

    Under our LPN Guidance, the second criterion we consider in 
assigning a listing priority is the immediacy of threats. This 
criterion is intended to ensure that the species facing actual, 
identifiable threats are given priority over those for which threats 
are only potential or that are intrinsically vulnerable but are not 
known to be presently facing such threats. We consider all of the 
threats to be imminent because we have information that the threats are 
identifiable and that the species are currently facing them across 
their entire range. These actual, identifiable threats are covered in 
greater detail in Factors A, D, and E of this finding. The majority of 
threats are ongoing and, therefore, imminent, although gravel mining is 
currently impacting only one of the populations. In addition to their 
current existence, we expect these threats to continue and likely 
intensify in the foreseeable future.
    The third criterion in our LPN guidance is intended to devote 
resources to those species representing highly distinctive or isolated 
gene pools as reflected by taxonomy. Eriogonum soredium and Lepidium 
ostleri are valid taxa at the species level and, therefore, receive a 
higher priority than subspecies, but a lower priority than species in a 
monotypic genus. Therefore, we assigned E. soredium and L. ostleri an 
LPN of 8.
    We will continue to monitor the threats to Eriogonum soredium and 
Lepidium ostleri and the species' status on an annual basis, and should 
the magnitude or the imminence of the threats change, we will revisit 
our assessment of the LPN.
    While we conclude that listing Eriogonum soredium and Lepidium 
ostleri is warranted, an immediate proposal to list this species is 
precluded by other higher priority listings, which we address in the 
Preclusion and Expeditious Progress section below. Because we have 
assigned Eriogonum soredium and Lepidium ostleri an LPN of 8, work on a 
proposed listing determination for Eriogonum soredium and Lepidium 
ostleri is precluded by work on higher priority listing actions with 
absolute statutory, court-ordered, or court-approved deadlines and 
final listing determinations for those species that were proposed for 
listing with funds from Fiscal Year (FY) 2010. This work includes all 
the actions listed in the tables included in the section on Preclusion 
and Expeditious Progress, below.

Species Information--Trifolium friscanum

Taxonomy and Species Description
    Trifolium friscanum is a dwarf mat-forming or tufted perennial herb 
in the legume family (Fabaceae). Plants have a taproot and thick woody 
stem. T. friscanum is up to 1.2 in (3 cm) tall and has silver hairy 
leaves composed of three leaflets (Welsh et al. 2008, p. 486). Its 
flowers resemble those of other clover species and are arranged in 
heads of four to nine reddish-purple flowers with pale wings (Welsh et 
al. 2008, p. 486). Flowering occurs from late May to June, followed by 
fruit set in June through July (Welsh et al. 2008, p. 486).
    Trifolium friscanum was originally described by Stanley Welsh as T. 
andersonii var. friscanum from specimens collected on Grampian Hill in 
the southern San Francisco Mountains in Beaver County, Utah (Welsh 
1978, p. 355). The variety was elevated to species level in 1993 (Welsh 
1993, p. 407). We accept the current taxonomy and consider T. friscanum 
to be a valid species and a listable entity under the ESA.
Distribution and Population Status
    Trifolium friscanum is a narrow endemic known from five small 
populations containing nine sites on private, SITLA, BLM, and USFS 
lands in Beaver and Millard Counties, Utah (Figure 4; Table 5; Kass 
1992c, pp. 4-5; Evenden 1998, pp. 6-7, Appendix C; Evenden 1999, pp. 2-
3; Miller 2010c, pp. 1, 4; Miller 2010e, pers. comm.; Roth 2010a, p. 
4). Populations are defined as groups of sites located in the same 
geographic vicinity. Sites are defined as occurrence records or 
locations recorded by one or more researcher over time within an 
individual population. Despite additional searches in the San Francisco 
Mountains and surrounding areas (including the Wah Wah Mountains, the 
Confusion Range, the Mountain Home Range, and the Tunnel Springs 
Mountains), no other populations are known to occur (Kass 1992c, pp. 4-
5; Evenden 1998, pp. 6-7, Appendix C; Evenden 1999, pp. 2-3; Miller 
2010c, pp. 1, 4; Miller 2010e, pers. comm.; Roth 2010a, p. 4).
    The five populations occur within three mountain ranges in 
southwestern Utah (see Figure 4 and Table 5). The two largest 
populations, the Grampian Hill and San Francisco Populations, occur on 
the southern tip on the San Francisco Mountains in Beaver County. East 
of the San Francisco Mountains are the Beaver Lake Mountains, where the 
Lime Mountain Population occurs on Lime Mountain. West and south of the 
San Francisco Mountains are the Wah Wah Mountains. Along the 
southeastern edge of the Wah Wah Mountains is the southernmost 
population, the Blue Mountain population, which occurs along the 
Beaver-Iron County boundary line on Blue Mountain. The Tunnel Springs 
Population occurs on Tunnel Springs Mountains in Millard County. The 
Tunnel Springs Mountains are west and north of the Wah Wah Mountains.
    Two of the five Trifolium friscanum populations overlap to some 
degree with the previously described Eriogonum soredium and Lepidium 
ostleri populations. The Grampian Hill populations of all three species 
occur on Grampian Hill on the southern tip of the San Francisco 
Mountains in the same habitat. The San Francisco population of T. 
friscanum overlaps with the Indian Queen populations of E. soredium and 
L. ostleri. The remaining three populations of T. friscanum--Blue 
Mountain, Lime Mountain, and Tunnel Springs--are located in nearby 
mountain ranges as described above.
BILLING CODE 4310-55-P

[[Page 10190]]

[GRAPHIC] [TIFF OMITTED] TR23FE11.003

BILLING CODE 4310-55-C

                             Table 5--Estimated Number of Trifolium friscanum Plants
  (Evenden 1998, Appendix C; Miller 2010a, pers. comm.; Miller 2010c, pp. 1, 4; 2010d, p. 1; Roth 2010a, p. 4).
----------------------------------------------------------------------------------------------------------------
                                                                        Estimated number of Trifolium friscanum
              Population                     Land ownership/sites                        plants
----------------------------------------------------------------------------------------------------------------
Blue Mountain.........................  SITLA (1 site)...............  250.

[[Page 10191]]

 
Grampian Hill.........................  Private (1 site).............  Many 1,000s.
San Francisco.........................  BLM (Copper Gulch) (1 site)..  1,000.
                                        Private (Cactus Mine) (1       300.
                                         site).
                                        Private (Indian Queen) (1      3,000.
                                         site).
Lime Mountain.........................  BLM (1 site).................  at least 125.
Tunnel Springs Mountains..............  BLM (1 site).................  500.
                                        USFS (2 sites)*..............  2,000.
    ESTIMATED TOTAL...................  .............................  13,000.
----------------------------------------------------------------------------------------------------------------
* Last surveyed in 1992. All other survey data from 2010.

    Trifolium friscanum populations extend about 40 mi (64 km) from the 
San Francisco Mountains and stretch across 650 mi\2\ (1,684 km\2\) 
(Figure 4). Within that area, the five populations are scattered in 
small, disjunct areas of occupied habitat (Figure 4; Table 5).
    The majority of plants (71 percent of the estimated populations) 
are located in the San Francisco and Grampian Hill populations (Miller 
2010g, Appendix B). Total occupied habitat for these two populations 
(four sites) is approximately 35 ac (14 ha), each site ranging between 
approximately 1 ac (0.4 ha) and 12 ac (5 ha) (Darnall et al. 2010, 
entire). The Blue Mountain population occupies an area of approximately 
0.33 ac (0.13 ha) (Darnall et al. 2010, entire). We do not have 
population estimates for the areas of occupied habitat for the Tunnel 
Springs sites (Tunnel Springs population) or the Lime Mountain 
population, but we assume the area of occupied habitat to be similar to 
or smaller than the San Francisco, Grampian Hill, and Blue Mountain 
populations, because these populations contain fewer than or similar 
numbers of plants as those estimated for the other sites (Table 5).
    The total number of Trifolium friscanum individuals in Table 5 was 
derived from observational counts or estimates. For the Grampian Hill 
population, the estimate was ``many thousands'' (Miller 2010a, pers. 
comm.). For the purpose of this finding, ``many thousands'' is 
interpreted as approximately 5,000 individuals. Four of the 9 sites 
contain 500 or fewer plants (Table 5).
    The population estimates were not based on actual counts of plants 
but on cursory observations with inherent observer biases. Similar to 
Eriogonum sorenium and Lepidium ostleri, the plants grow in dense mat-
forming clusters, making it difficult to determine the number of 
individuals within a cluster. Because individual plants are difficult 
to distinguish, we do not believe that the variation in population 
estimates reflects variation in population sizes, but is rather an 
artifact in survey effort and methods used. Many of the sites occur on 
private lands where access is restricted, so population counts are 
estimated from observations.
    Accordingly, the available population estimates are highly variable 
and probably not accurate. During the 1990s, population estimates 
ranged from 3,500 individuals (Evenden 1998, Appendix C) to 
approximately 6,000 individuals (Kass 1992c, p. 8). In 2010, the total 
number of plants was estimated at roughly 13,000 (Table 5; Miller 
2010a, pers. comm.; Miller 2010c, pp. 1, 4; Miller 2010d, p. 1; Roth 
2010a, p. 4). Thus, we do not have accurate population estimates or 
trends for this species.
Habitat
    Trifolium friscanum is a narrow endemic restricted to soils derived 
from volcanic gravels, Ordovician limestone, and dolomite outcrops. 
Soils are shallow, with gravels, rocks, and boulders on the surface 
(Kass 1992c, p. 3; Miller 2010d, p. 1).
    In the southern San Francisco Mountains, where the majority of 
plants are located, there are 845 ac (342 ha) of Ordovician limestone 
and 719 ac (291 ha) of dolomite outcrops (Darnall et al. 2010, entire). 
Ordovician limestone is rare within a 50-mi (80-km) radius of the San 
Francisco Mountains, but dolomite outcrops are common in the Wah Wah 
Mountain Range to the west (Miller 2010g, Appendix F). We have no 
information on the extent of volcanic gravels in the area. As 
previously described (see Distribution and Population Status), we are 
not aware of any additional populations of the species, despite 
additional potentially suitable habitats.
    We do not know if there are other limiting factors associated with 
the limestone and dolomite formations that restrict the habitat use and 
distribution of the species; the species occupies only a fraction of 
the available habitat. The two largest populations--Grampian Hill and 
San Francisco--occupy an estimated 35 ac (14 ha) (2.3 percent) of the 
available limestone and dolomite outcrops (Darnall et al. 2010, 
entire). We do not have occupied habitat area totals for the remaining 
three populations, but we believe they are smaller, based on field 
evaluations and the lower number of individuals in these populations 
(Kass 1992c, p. 3; Miller 2010d, p. 1; Roth 2010a, pp. 1-2).
    Trifolium friscanum is typically found within sparsely vegetated 
pinion-juniper-sagebrush communities between 5,640 and 8,440 ft (1,720-
2,573 m) in elevation. Associated species include Ephedra spp. (Mormon 
tea), Gutierrezia sarothrae (snakeweed), Cercocarpus intricatus (dwarf 
mountain-mahogany), and Petradoria pumila (rock goldenrod). Associated 
rare species in the southern San Francisco Mountains include Eriogonum 
soredium and Lepidium ostleri, which generally grow on the same 
substrate in similar but more open habitats adjacent to T. friscanum.
Life History
    No information is available on the life history of this species.

Summary of Information Pertaining to the Five Factors--Trifolium 
friscanum

    In making our 12-month finding on the petition, we considered and 
evaluated the best available scientific and commercial information 
pertaining to Trifolium friscanum in relation to the five factors 
provided in section 4(a)(1) of the ESA (see the full description of 
these five factors in the Summary of Information Pertaining to the Five 
Factors--Astragalus hamiltonii, above).

Factor A. The Present or Threatened Destruction, Modification, or 
Curtailment of Its Habitat or Range

    The following factors may affect the habitat or range of Trifolium 
friscanum:

[[Page 10192]]

(1) Livestock grazing, (2) recreational activities, (3) mining, and (4) 
nonnative invasive species.
(1) Livestock Grazing
    Potential impacts of livestock grazing to plants are discussed 
above in the Livestock Grazing section under Factor A for Astragalus 
hamiltonii.
    All Trifolium friscanum populations on BLM lands are located on 
active grazing allotments (Galbraith 2010, pers. comm.). Adjacent 
habitats on SITLA and private lands are subject to the same grazing 
practices as the allotted BLM land if the habitats are not fenced 
(Galbraith 2010, pers. comm.). The SITLA and private lands are only 
partially fenced in these areas; thus we can assume that grazing 
occurs. The USFS sites of the Tunnel Springs population are not grazed 
(Kitchen 2010, pers. comm.).
    The Trifolium friscanum population on BLM lands in the Tunnel 
Springs Mountains was likely impacted by the construction of an 
allotment boundary fence 10 years ago (Evenden 1999, p. 7; Roth 2010a, 
p. 2). The fence runs along a ridge and through approximately 500 ft 
(150 m) of T. friscanum habitat (Roth 2010b, p.1). The construction of 
the fence may have impacted approximately 10 percent of the species' 
habitat in the area (Roth 2010b, p.1). Livestock and wildlife trailing 
occur along the fence, resulting in trampling of individual plants and 
soil compaction (Roth 2010a, p. 2). No plants occur within 100 ft (30 
m) of either side of the fence (Roth 2010a, p. 2).
    Although much of the species' habitat is accessible to livestock, 
we are not aware of any other disturbances or loss of plants from 
grazing (Kass 1992, entire; Evenden 1998, entire, Evenden 1999, entire; 
Pontarolo 2009, pers. comm.; Miller 2010f, pers. comm.; Roth 2010a, p. 
3). Available information suggests that livestock grazing is not 
occurring at a level that is impacting the species (Pontarolo 2009, 
pers. comm.; Miller 2010f, pers. comm.; Roth 2010a, p. 3). Therefore, 
we have no information suggesting that grazing impacts the species now 
or will impact the species in the foreseeable future at a level that 
threatens Trifolium friscanum.
(2) Recreational Activities
    Potential impacts of recreational activities to plants are 
discussed above in the Recreational Activities Section, Factor A, for 
Astragalus hamiltonii. Because we know that OHV use is widespread 
across the southwestern landscape, we analyzed its occurrence in 
Triolium friscanum's habitat for this finding.
    Access to the majority of occupied habitat on private lands is 
closed to all vehicles, including OHVs (Miller 2010g, p. 5). There are 
no known impacts of OHV use in Trifolium friscanum's occupied habitat 
on private lands (Miller 2010f, pers. comm.; Roth 2010a, pp. 1-2). The 
OHV use also does not appear to impact T. friscanum's habitat on SITLA, 
BLM, or USFS lands (Pontarolo 2009, pers. comm.; 2010, pers. comm.; 
Miller 2010f, pers. comm.; Roth 2010a, pp. 1-2). Therefore, we do not 
believe that recreational activities threaten T. friscanum now, nor do 
we anticipate that these activities will become a threat in the 
foreseeable future.
(3) Mining
    As previously described (see Distribution and Population Status), 
Trifolium friscanum occurs in five population areas: Blue Mountain, 
Grampian Hill, San Francisco, Lime Mountain, and Tunnel Springs 
Mountains. For purposes of the following analysis, it is important to 
note that the Grampian Hill and San Francisco populations occur in the 
southern San Francisco Mountains in the same vicinity and habitat as 
Eriogonum soredium and Lepidium ostleri. The other three populations 
are located in nearby mountain ranges.
    The San Francisco Mountains have an extensive history of mining of 
precious metals and limestone gravel (Table 6; Evenden 1998, p. 3). We 
described this mining history, the likelihood of future mining 
activities, and effects to the species under Eriogonum soredium and 
Lepidium ostleri, Factor A, Mining. This analysis applies to the 
Grampian Hill and San Francisco populations of Trifolium friscanum, 
because the three species co-occur (see Distribution and Population 
Status). In addition, we evaluated mining activity and its impacts to 
the remaining three populations of T. friscanum.
    To review, precious metal mining in the southern San Francisco 
Mountains is likely to impact the Grampian Hill and San Francisco 
populations of Trifolium friscanum (Table 6). The Grampian Hill 
population is located in the area of the King David Mine, which is part 
of the historical Horn Silver Mine. The San Francisco population (which 
overlaps the Indian Queen population of Eriogonum soredium and Lepidium 
ostleri) is in the vicinity of mine shafts near the Cactus Mine, an 
historical copper mine (see E. soredium and L. ostleri, Factor A, 
Mining). Although large-scale precious metal mining in the area ceased 
decades ago, we believe mining is likely to occur again in the 
foreseeable future due to patent rights and ongoing exploration for 
silver, zinc, and copper deposits--including recent exploration 
activities at the Horn Silver Mine (see E. soredium and L. ostleri, 
Factor A, Mining). Precious metal mining in the vicinity of the 
Grampian Hill and San Francisco populations is of concern because these 
populations comprise the species' largest known populations, containing 
the vast majority of known individuals (9,300 individuals, or 71 
percent of the species' estimated total population) (Table 5).
    The Lime Mountain population has experienced precious metal mining 
activity in the past (Table 6; Miller 2010h, pp. 6-7). The last mining 
activity occurred in the early 1980s. We do not anticipate additional 
mining, due to the small amounts of minerals that were extracted 
(Miller 2010h, p. 7). We are not aware of precious metal mining 
activities in the vicinity of the Blue Mountain or Tunnel Springs 
populations.

                        Table 6--Mining activities in the habitat of Trifolium friscanum
----------------------------------------------------------------------------------------------------------------
                                                                    Mining Activity
              Population              --------------------------------------------------------------------------
                                              Historical                Current                   Future
----------------------------------------------------------------------------------------------------------------
Blue Mountain........................  gravel quarrying.......  active.................  gravel quarrying.
Grampian Hill........................  silver, lead, copper,    none...................  silver, lead, copper,
                                        zinc (Horn Silver                                 zinc, landscape gravel
                                        Mine).                                            quarrying.
San Francisco........................  silver, lead, copper,    active.................  silver, lead, copper,
                                        zinc, gravel quarrying                            zinc, landscape gravel
                                        (Cactus Mine).                                    quarrying.

[[Page 10193]]

 
Lime Mountain........................  silver, lead, copper,    none...................  unknown.
                                        zinc, native gold,
                                        iron (Skylark,
                                        Independence & Galena
                                        Mines).
Tunnel Springs Mountains.............  unknown................  none...................  unknown.
----------------------------------------------------------------------------------------------------------------

    Gravel mining is known to occur within the range of Trifolium 
friscanum, particularly in the San Francisco Mountains and Wah Wah 
Mountains. Impacts to T. friscanum from gravel mining in the southern 
San Francisco Mountains is similar to those analyzed for Eriogonum 
soredium and Lepidium ostleri, because of their co-occurrence (see E. 
soredium and L. ostleri, Factor A, Mining, above).
    Gravel mining in the southern San Francisco Mountains is likely to 
impact the San Francisco population of T. friscanum and possibly the 
Grampian Hill population (Table 6). We estimate that 19 ac (8 ha) of 
suitable habitat is disturbed by gravel mining activities near the San 
Francisco population of Trifolium friscanum. Two quarries are located 
within 1,000 ft (300 m) of two sites (Cactus Mine and Copper Gulch) of 
the San Francisco population of T. friscanum. Based on habitat 
similarities and proximity, we believe the plant may have occupied 
these areas prior to the mining activity. Gravel pits in this area are 
considered active because they are not reclaimed--given their close 
proximity to known T. friscanum plants, these gravel pits could impact 
the remaining occupied habitat of the species through additional 
quarrying activities (i.e., removal of the entire substrate) or when 
roads and other infrastructure are constructed. The San Francisco 
population currently occupies only 15 ac (6 ha) of habitat, distributed 
in three sites (Copper Gulch, Cactus Mine, and Indian Queen) (Table 5; 
Darnall et al. 2010, entire).
    Gravel mining also may impact the Grampian Hill population of 
Trifolium friscanum in the future. Although gravel mining is not 
actively occurring at Grampian Hill, gravel pits exist within 1 mi (1.6 
km) of this T. friscanum population--near the Cupric Mine (see E. 
soredium and L. ostleri, Factor A, Mining, above). We do not know if 
gravel mining will definitely occur at the Grampian Hill population. 
However, mining operations are expected to either expand from the 
vicinity of the Cupric Mine or be moved to a new location within the 
species' habitat in the near future (Munson 2010, pers. comm.). Due to 
the limited extent of the Ordovician limestone deposits across the 
landscape (see Habitat), it is plausible that mining activities could 
occur at the Grampian Hill population. Even if gravel mining does not 
occur at the Grampian Hill population, we previously established that 
this population is likely to be impacted by precious metal mining.
    A similar overlap in habitat types and gravel quarrying (Table 6) 
occurs for this species in the Blue Mountain population. The Blue 
Mountain population, which is less than 1 ac (0.4 ha) in size, is 
located on SITLA lands within a couple hundred feet (meters) of a 
gravel pit (Evenden 1998, p. 9; Roth 2010a, p. 4). This mine is not 
reclaimed and, therefore, is considered active (Darnall et al. 2010, 
entire). Therefore, we assume that continued gravel mining will 
ultimately impact this population if it has not already occurred. The 
need for gravel sources is expected to increase, because an increasing 
human population growth (U.S. Census Bureau 2010b, entire; Utah GOPB 
2010, p. 48) will result in the need for increased road construction 
and maintenance in the future. Although the gravel in the Blue Mountain 
is mined for road construction projects, the effects analysis under E. 
soredium and L. ostleri (see Factor A, Mining) is relevant; i.e., 
mining for gravel will lead to the degradation and loss of suitable 
habitat for Trifolium friscanum.
    As previously discussed (see Eriogonum soredium and Lepidium 
ostleri, Factor A, Mining, above), construction sand, gravel, and 
crushed stone together rank as the second most valuable commodity 
produced among industrial minerals in Utah (Bon and Krahulec 2009, p. 
5). Gravel, stone, and rock are generally mined for local and regional 
distribution due to the high cost of transport. The quarries in the San 
Francisco Mountains are the closest crushed limestone quarries to 
Washington County, one of the fastest growing counties in Utah (see E. 
soredium and L. ostleri, Factor A). In general, there has been a net 
loss of local sand and gravel supply pits in the Washington County area 
due to ongoing urban development and the lack of available gravel pit 
operations on surrounding Federal lands (Blackett and Tripp 1999, p. 
33). Thus, the Blue Mountain population area could become a primary 
source of gravel for Washington County and other nearby communities, 
especially because the pit's location on SITLA lands limits the need 
for environmental regulations. Overall, it is likely that an increasing 
human population growth in Washington County (U.S. Census Bureau 2010b, 
entire; Utah GOPB 2010, p. 48) will result in an increased demand for 
the limestone and gravel resources at and nearby known populations of 
T. friscanum.
    To summarize, mining throughout large portions of Trifolium 
friscanum's range has impacted available habitat. Three of the five 
known populations are located at historical precious metal mines or 
gravel mines on private and SITLA lands (Table 5; Table 6; see Factor 
D). Two of these populations (San Francisco and Grampian Hill) comprise 
the vast majority (71 percent) of the known estimated population of T. 
friscanum (Table 5). Precious metal mining is likely to impact 
populations of T. friscanum in the foreseeable future, particularly in 
the vicinity of the large Grampian Hill and San Francisco populations. 
Gravel mining is expected to increase in the future in response to 
increased population growth and limited availability of active gravel 
pits in nearby Washington County (see E. soredium and L. ostleri, 
Factor A). Available information suggests that three of five 
populations will be significantly impacted by either precious metal or 
gravel mining in the foreseeable future (see E. soredium and L. 
ostleri, Factor A, Mining). Therefore, we have determined that mining 
is a threat to T. friscanum now and in the foreseeable future.
(4) Nonnative Invasive Species
    Potential impacts of nonnative invasive species to native plants 
and their habitat are discussed above in Astragalus hamiltonii, Factor 
A, Nonnative Invasive Species. The annual nonnative invasive grass, 
Bromus tectorum, is considered the most

[[Page 10194]]

ubiquitous invasive species in the Intermountain West due to its 
ability to rapidly invade native dryland ecosystems and outcompete 
native plant species (Mack 1981, p. 145; Mack and Pyke 1983, p. 88; 
Thill et al. 1984, p. 10).
    Bromus tectorum occurs in the habitat and vicinity of the Grampian 
Hill and San Francisco Trifolium friscanum populations, which also is 
where the majority of plants occur (Table 5; Miller 2010c, pp. 2-5; 
Roth 2010a, p. 1). We do not know whether B. tectorum occurs in the 
other three populations, but given the ubiquitous distribution of B. 
tectorum in the Intermountain West, we expect it occurs in the vicinity 
of all populations (Novack and Mack, 2001, p. 115).
    Surface disturbances increase the occurrence and densities of B. 
tectorum (see Eriogonum soredium and Lepidium ostleri, Factor A, 
Nonnative Invasive Species; Mack 1981, p. 145). As previously 
described, increased mining activities and associated surface 
disturbances are expected to occur in and adjacent to the occupied 
habitat for T. friscanum in the San Francisco and Blue Mountains (see 
Mining, above), consequently encouraging B. tectorum to expand into the 
species' habitat.
    Invasions of annual nonnative species, such as Bromus tectorum, are 
well documented to contribute to increased fire frequencies (Brooks and 
Pyke 2002, p. 5; Grace et al. 2002, p. 43; Brooks et al. 2003, pp. 4, 
13, 15). The risk of fire is expected to increase from 46 to 100 
percent when the cover of B. tectorum increases from 12 to 45 percent 
or more (Link et al. 2006, p. 116). In the absence of exotic species, 
it is generally estimated that fire return intervals in xeric sagebrush 
communities range from 100 to 350 years (Baker 2006, p. 181). In some 
areas of the Great Basin (Snake River Plain), fire return intervals due 
to B. tectorum invasion are now between 3 and 5 years (Whisenant 1990, 
p. 4). Most plant species occurring within a sagebrush ecosystem are 
not expected to be adapted to frequent fires, as evidenced in the lack 
of evolutionary adaptations found in other shrub-dominated fire-adapted 
ecosystems like chaparral. Examples of such adaptation would include 
re-sprouting and heat-stimulated seed germination (Baker, in press, p. 
17).
    In the absence of annual nonnative species, T. friscanum grows in 
sparsely vegetated communities that are unlikely to carry fires (see 
Habitat section). Thus, T. friscanum is unlikely to be adapted to fire 
and, therefore, unlikely to persist through a fire. Therefore, the 
potential expansion of invasive species and associated fire is a threat 
to the species, especially when considering the limited distribution of 
the species and the high potential of stochastic extinctions (as 
discussed in the Small Population Size, Factor E, below). As described 
in the Distribution section, the majority of plants are located within 
the Grampian Hill and San Francisco populations, where occurrences of 
B. tectorum are documented. Occupied habitat in these populations 
ranges from 1 to 12 ac (0.4 to 5 ha).
    In summary, Bromus tectorum occurs in the two largest Trifolium 
friscanum populations (Grampian Hill and San Francisco populations, 
Table 5). Given the ability of B. tectorum to rapidly invade dryland 
ecosystems (Mack 1981, p. 145; Mack and Pyke, 1983, p. 88; Thill et al. 
1984, p. 10), we expect it to increase in the future in response to 
surface disturbance from increased mining activities and global climate 
change (see the Climate Change and Drought section under Factor E for 
Astragalus hamiltonii). An increase in nonnative species is expected to 
increase the frequency of fires in T. friscanum's habitat. Therefore, 
we determine that nonnative invasive species are a threat to two of 
five populations of T. frsicanum and the majority of individuals now, 
and may impact all populations in the foreseeable future when evaluated 
cumulatively with mining activities (and associated surface 
disturbances), climate change, and fire.
Summary of Factor A
    At this time, based on best available information, we do not 
believe that grazing or recreational activities significantly threaten 
Trifolium friscanum now or in the foreseeable future. However, we 
determine that mining and nonnative invasive species are threats to T. 
friscanum.
    Mining activities impacted Trifolium friscanum habitat in the past 
and continue to be a threat to the species and its habitat throughout 
large portions of its range. Two of the five populations and the 
majority of individuals are located on lands with an extensive history 
of precious metal mining; ongoing exploration activities indicate that 
precious metal mining is likely to threaten the species in the 
foreseeable future. The main threat to the majority of T. friscanum 
plants is gravel mining (Table 6). Three of the five populations are 
located in the vicinity of gravel pits that are mined for road and 
landscaping gravel. The three populations located in the vicinity of 
gravel mines contain the majority of plants and may be mined for gravel 
in the future (Table 6). We anticipate an increase in the demand for 
precious metals and landscape rock based on the economic outlook for 
these commodities, regional availability, and the proximity of these 
gravel mines to a rapidly expanding urban area and, therefore, an 
increase in impacts to T. friscanum.
    Bromus tectorum is documented to occur in the two largest of the 
five populations of Trifolium friscanum. The threat of fire caused by 
annual nonnative species invasions is exacerbated by mining activities 
and global climate change (see the Climate Change and Drought section 
under Factor E). Small population sizes and extremely limited 
distribution of this species make it especially vulnerable to 
stochastic extinction events, including mining activities and wildfires 
caused by increased invasions of nonnative species (see the Small 
Population Size section under Factor E).
    Therefore, we find that Trifolium friscanum is threatened by the 
present or threatened destruction, modification, or curtailment of the 
species' habitat or range, now and in the foreseeable future, based on 
impacts from mining activities and nonnative invasive species.

Factor B. Overutilization for Commercial, Recreational, Scientific, or 
Educational Purposes

    Trifolium friscanum is not a plant of horticultural interest. We 
are not aware of any overutilization or collection of T. friscanum. 
Therefore, overutilization for commercial, recreational, scientific, or 
educational purposes does not appear to pose a significant threat to 
the species now nor is it likely to become a threat in the foreseeable 
future.

Factor C. Disease or Predation

    Disease and herbivory on the species are unknown. We do not have 
any information indicating that disease is impacting Trifolium 
friscanum. We also do not have any information indicating that 
herbivory is occurring from livestock (see the Livestock Grazing 
section under Factor A), wildlife, or insects (Kass 1992c, p. 10; 
Evenden 1998, entire; Evenden 1999, entire; Miller 2010a, p. 1; Miller 
2010c, entire; Roth 2010a, entire). Thus, we do not consider disease or 
predation to be threats to this species.

Factor D. The Inadequacy of Existing Regulatory Mechanisms

    There are no endangered species laws protecting plants on private, 
State, or Tribal lands in Utah. The majority of individual plants are 
located on SITLA

[[Page 10195]]

or private lands (Table 5). Trifolium friscanum is listed as a bureau-
sensitive plant for the BLM. Limited policy-level protection by the BLM 
is afforded through the Special Status Species Management Policy Manual 
 6840, which forms the basis for special status species 
management on BLM lands (BLM 2008e, entire). The two sites on USFS 
lands are located within the Desert Experimental Range in the Tunnel 
Springs Mountains (Tunnel Springs population) and appear to be secure, 
although the population has not been visited since 1992 (Kass 1992c, p. 
11; Evenden 1998, Appendix C; Evenden 1999, p. 3).
    This species is predominantly located on private or SITLA lands 
(Table 5), where it is threatened by mining-related activities (see 
Factor A). There are limited regulatory mechanisms in place that may 
protect Trifolium friscanum from mining on private or State lands. As 
described under Eriogonum soredium and Lepidium ostleri, Factor D, 
State environmental impact assessments are required for large mining 
operations for all mineral exploration, development, and extraction, 
including gravel pits and precious metal mining (UDOGM 2010b, p.1; 
Baker 2010, pers. comm.). T. friscanum is not State listed, but it is 
on the BLM sensitive species list. If UDOGM is made aware of impacts to 
these species, they could consider minimizing and mitigating impacts; 
however, there is no requirement to address species that are not 
federally listed in the mine permitting process (Baker 2010, pers. 
comm.).
    The existing mining activities (see Factor A, Mining) are under the 
5-ac (2-ha) regulatory threshold and, therefore, not subject to 
permitting laws (Munson 2010, pers. comm.). A few of the gravel mine 
pits almost exceed the 5-ac (2-ha) limit, and the operators may need to 
apply for permits (Munson 2010, pers. comm.); however, they also could 
choose to begin new gravel pits, or reclaim portions of the existing 
pits to remain below the 5-ac (2-ha) limit (Munson 2010, pers. comm.).
    In summary, the existing regulatory mechanisms are not adequate to 
protect T. friscanum from becoming threatened or endangered by precious 
metal or gravel mining on SITLA and private lands. The active gravel 
pits are below the 5-ac (2-ha) threshold that would automatically 
trigger regulatory environmental impact assessments. Even if an 
environmental impact assessment is completed for any of the mines, the 
existing mining laws only recommend, and do not mandate, the species' 
protection or mitigation. Thus, we find that the inadequacy of existing 
mechanisms to regulate mining activities on private and State lands is 
a threat to three of five populations and the majority of individuals, 
and thus to T. friscanum now and into the foreseeable future.

Factor E. Other Natural or Manmade Factors Affecting Its Continued 
Existence

    Natural and manmade threats to Trifolium friscanum's survival 
include: (1) Small population size and (2) climate change and drought.
(1) Small Population Size
    General potential impacts of small population sizes in plants are 
discussed above in the Small Population Size section under Factor E for 
Astragalus hamiltonii.
    As previously discussed (see Distribution and Population Status, 
above), the entire species' range is restricted to highly specialized 
habitat niches, distributed in 5 populations (and 9 sites) with a total 
population estimate of 13,000 plants. Four of the 9 sites contain 500 
or fewer individuals (Table 5). Only a fraction of the entire species' 
range is occupied habitat. The majority of plants are located in two 
populations containing four sites of occupied habitat, ranging from an 
estimated 1 ac (0.4 ha) to a maximum of 12 ac (5 ha) (Darnall et al. 
2010, entire; Miller 2010g, Appendix B).
    Despite the overall lack of information on the population ecology 
of Trifolium friscanum, we know that small populations are at an 
increased risk of extinction due to the potential for inbreeding 
depression, loss of genetic diversity, and lower sexual reproduction 
rates (Ellstrand and Elam 1993, entire; Wilcock and Neiland 2002, p. 
275). No information is available on the population genetics, 
pollination, or reproductive effort and success of T. friscanum. 
However, the small areas of occupation and the narrow overall range of 
the species make it highly susceptible to stochastic extinction events 
and the effects of inbreeding depression.
    Mining or a single random event, such as a wildfire from invasive 
species (see Factor A, Nonnative Invasive Species), could extirpate an 
entire or at least a substantial portion of a population, given the 
small areas of occupied habitat. Species with limited ranges and 
restricted habitat requirements also are more vulnerable to the effects 
of global climate change (see Climate Change and Drought, below) (IPCC 
2002, p. 22; Jump and Penuelas 2005, p. 1016; Machinski et al. 2006, p. 
226; Krause 2010, p. 79). Overall, we consider small population size an 
intrinsic vulnerability to Trifolium friscanum, which may not rise to 
the level of a threat on its own. However, the small population sizes 
rise to the level of a threat because of the combined effects of having 
only five highly localized small populations with the effects of global 
climate change (see below) and the potential for stochastic extinction 
events such as mining, and fire induced by invasive species (see Factor 
A). Therefore, we consider small localized population size, in 
combination with mining, invasive species, and climate change, to be a 
threat to the species now and in the foreseeable future.
(2) Climate Change and Drought
    Potential impacts of climate change and drought to the geographic 
area are characterized in the Climate Change and Drought section under 
Factor E for Astragalus hamiltonii. As discussed in the Small 
Population Size section above, Trifolium friscanum has a limited 
distribution and populations are localized and small. In addition, 
these populations are restricted to very specific soil types. Global 
climate change exacerbates the risk of extinction for species that are 
already vulnerable due to low population numbers and restricted habitat 
requirements (see Climate Change and Drought, Factor E for Astragalus 
hamiltonii, above).
    Predicted changes in climatic conditions include increases in 
temperature, decreases in rainfall, and increases in atmospheric carbon 
dioxide in the American Southwest (Walther et al. 2002, p. 389; IPCC 
2007, p. 48; Karl et al. 2009, p. 129). Although we have no information 
on how Trifolium friscanum will respond to effects related to climate 
change, persistent or prolonged drought conditions are likely to reduce 
the frequency and duration of flowering and germination events, lower 
the recruitment of individual plants, compromise the viability of 
populations, and impact pollinator availability (Tilman and El Haddi 
1992, p. 263; Harrison 2001, p. 78). The smallest change in 
environmental factors, especially precipitation, plays a decisive role 
in plant survival in arid regions (Herbel et al. 1972, p. 1084).
    Drought conditions led to a noticeable decline in survival, vigor, 
and reproductive output of other rare and endangered plants in the 
Southwest during the drought years of 2001 through 2004 (Anderton 2002, 
p. 1; Van Buren and Harper 2002, p. 3; Van Buren and Harper 2004, 
entire; Hughes 2005, entire; Clark and Clark 2007, p. 6; Roth 2008a, 
entire; Roth 2008b, pp. 3-4). Similar responses are anticipated to

[[Page 10196]]

adversely affect the long-term persistence of T. friscanum.
    Climate change is expected to increase levels of carbon dioxide 
(Walther et al. 2002, p. 389; IPCC 2007, p. 48; Karl et al. 2009, p. 
129). Elevated levels of carbon dioxide lead to increased invasive 
annual plant biomass, invasive seed production, and pest outbreaks 
(Smith et al. 2000, p. 80-81; IPCC 2002, pp. 18, 32; Ziska et al. 2005, 
p. 1328), and will put additional stressors on rare plants already 
suffering from the effects of elevated temperatures and drought.
    The actual extent to which climate change itself will impact 
Trifolium friscanum is unclear, mostly because we do not have long-term 
demographic information that allows us to predict the species' response 
to changes in environmental conditions, including prolonged drought. 
However, as previously described, the species is threatened by mining 
activities (see Mining, Factor A, above), which will likely result in 
the loss of large numbers of individuals or even entire populations. 
Increased surface disturbances associated with mining activities also 
will likely increase the extent and densities of nonnative invasive 
species and, with these, the frequencies of fires (see Nonnative 
Invasive Species, Factor A, above). The cumulative effects of the 
potential reduction in population numbers and habitat loss (of already 
small populations) associated with mining and increased invasive 
species (and fire) are likely to increase the risk of the species being 
impacted by changes in climate.
    In summary, we find it difficult to analyze the potential effects 
of global climate change on Trifolium friscanum in the absence of 
demographic trend data for the species which would allow us to analyze 
how the species responds to climate change through time. However, the 
cumulative effects posed by the threats of mining, nonnative species 
and small population size may exacerbate the effects of climate change 
on T. friscanum in the future. However, at this time, we believe that 
the state of knowledge concerning the localized effects of climate 
change within the habitat occupied by T. friscanum is too speculative 
to determine whether climate change is a threat to this species in the 
foreseeable future. We will continue to assess the potential of climate 
change to threaten the species as better scientific information becomes 
available.
Summary of Factor E
    We assessed the potential risks of small population size, climate 
change, and drought to Trifolium friscanum populations. T. friscanum 
has a highly restricted distribution and is known from five small, 
localized populations. Even in the absence of information on genetic 
diversity, inbreeding depression, and reproductive effort, a random 
stochastic event could impact a significant portion of a population. 
Small populations that are restricted by habitat requirements are also 
more vulnerable to the effects of climate change, such as prolonged 
droughts and increased fire frequencies.
    While naturally occurring droughts are not likely to impact the 
long-term persistence of the species, an increase in periodic prolonged 
droughts due to climate change is likely to impact the species across 
its entire range in the future. Global climate change, particularly 
when assessed cumulatively with small population size and threats from 
mining activities, is expected to increase the density of invasive 
annual grasses, which are already present in the habitat of Trifolium 
friscanum within the populations that contain the majority of the 
plants (see Factor A). Increased nonnative species in the habitat of T. 
friscanum can increase fire frequency and severity. Because T. 
friscanum is not likely adapted to persist through fires, wildfires can 
have a significant impact on these small populations.
    Although small population size and climate change make the species 
intrinsically more vulnerable, we are uncertain whether they would rise 
to the level of threat by themselves. However, when combined with the 
threats listed under Factor A, we believe that small population size is 
likely to rise to the level of threat in the foreseeable future. At 
this time, we are uncertain of the degree to which climate change 
constitutes a threat to the species.

Finding

    As required by the ESA, we conducted a review of the status of the 
species and considered the five factors in assessing whether Trifolium 
friscanum is endangered or threatened throughout all or a significant 
portion of its range. We examined the best scientific and commercial 
information available regarding the past, present, and future threats 
faced by T. friscanum. We reviewed the petition, information available 
in our files, as well as other available published and unpublished 
information, and we consulted with species experts and other Federal 
and State agencies.
    This status review identified threats to the species attributable 
to Factors A, D, and E. The primary threat to the species is habitat 
destruction from precious metal and gravel mining on private and SITLA 
lands (Factor A). The largest populations containing the majority of 
Trifolium friscanum plants are located on private lands with active 
mining claims. These populations were likely impacted by historical 
precious metal mining. Another population is located on SITLA lands in 
the immediate vicinity of a gravel pit. We expect an increase in 
precious metal and gravel mining in the foreseeable future, with the 
associated loss and fragmentation of T. friscanum populations.
    Bromus tectorum occurs in the vicinity of the two largest 
populations of the five known Trifolium friscanum populations. It is a 
highly invasive species and is expected to increase in areas where 
surface disturbance such as mining occurs. As previously discussed, the 
species occurs in the vicinity of gravel and precious metal mines. 
Mines inherently cause surface disturbances from excavation activities 
and the construction of roads and other infrastructure. Global climate 
change is expected to increase drought conditions in the Southwest and 
increase the spread of nonnative invasive species. The biggest concern 
associated with the increase in invasive species is the threat of 
increased wildfire (Factor A), particularly when considering the small 
population sizes and small occupied habitat acreages associated with 
the species.
    The magnitude of the biological threats posed by the small 
population size and limited species range are not well understood due 
to the lack of information available on the ecology of Trifolium 
friscanum. Future studies may provide us with a more thorough 
understanding of threats posed by pollinator limitation, inbreeding 
depression, and the potential lack of genetic diversity over the 
species' range. Even without detailed knowledge on how small population 
sizes are impacting the biology and ecology of T. friscanum, the small 
areas of occupied habitat make the species highly vulnerable to habitat 
destruction through mining-related activities as well as random 
extinction events, including fires and the effects of global climate 
change (Factor E).
    The existing regulatory mechanisms are not adequate to protect 
Trifolium friscanum from the primary threat of mining, particularly 
because the

[[Page 10197]]

majority of individuals are located on private lands (Factor D). The 
inadequacy of regulatory mechanisms (Factor D) on private and State 
lands, combined with the high economic and commercial value of much of 
the substrate this species depends on, poses a serious threat to T. 
friscanum. A large portion of the species' individuals have the 
potential to be extirpated by mining activities in the foreseeable 
future (Factor A; Table 6). Ongoing mining in the habitat of T. 
friscanum has the potential to extirpate three of the five populations 
in the foreseeable future, two of which contain the majority of plants 
(Factor A, Table 5).
    On the basis of the best scientific and commercial information 
available, we find that the petitioned action to list Trifolium 
friscanum as endangered or threatened is warranted. We will make a 
determination on the status of the species as endangered or threatened 
when we do a proposed listing determination. However, as explained in 
more detail below, an immediate proposal of a regulation implementing 
this action is precluded by higher priority listing actions, and 
progress is being made to add or remove qualified species from the 
Lists of Endangered and Threatened Wildlife and Plants.
    We reviewed the available information to determine if the existing 
and foreseeable threats render the species at risk of extinction now 
such that issuing an emergency regulation temporarily listing the 
species under section 4(b)(7) of the ESA is warranted. We determined 
that issuing an emergency regulation temporarily listing the species is 
not warranted at this time because there is no emergency posing a 
significant risk to the well being of Trifolium friscanum. We do not 
believe that any of the potential threats are of such great immediacy 
and severity that would threaten all of the known populations with the 
imminent risk of extinction. However, if at any time we determine that 
issuing an emergency regulation temporarily listing Trifolium friscanum 
is warranted, we will initiate this action at that time.

Listing Priority Number

    Pursuant to our guidelines, titled ``Endangered and Threatened 
Species Listing and Recovery Priority Guidelines'' (described above), 
we have assigned Trifolium friscanum a Listing Priority Number (LPN) of 
8, based on our finding that the species faces threats that are of 
moderate magnitude and are imminent. These threats include the present 
or threatened destruction, modification, or curtailment of its habitat, 
the inadequacy of existing regulatory mechanisms, and other natural or 
manmade factors affecting its continued existence. These threats are 
ongoing and, in some cases (such as nonnative species), are considered 
irreversible because large-scale invasions cannot be recovered to a 
native functioning ecosystem. Our rationale for assigning T. friscanum 
an LPN of 8 is outlined below.
    Under the Service's LPN guidance, the magnitude of threat is the 
first criterion we look at when establishing a listing priority. The 
guidance indicates that species with the highest magnitude of threat 
are those species facing the greatest threats to their continued 
existence. These species receive the highest listing priority. We 
consider the magnitude of Factor A moderate. While current mining 
activities are ongoing in the habitat of T. friscanum, they are not 
ongoing in the immediate vicinity of any of the populations. Mining in 
the habitat of these populations is expected to increase the density of 
B. tectorum, thereby facilitating the spread of fire. B. tectorum 
occurs in two of the five populations, which also contain the largest 
number of individuals. We have no documentation on the density of B. 
tectorum within these populations but we are expecting it to increase 
in the future.
    We consider the magnitude of Factor D to be moderate. Three of the 
five populations are located on private or SITLA lands. The majority of 
individuals are located on private lands with active patented mining 
claims. Existing regulatory mechanisms do not adequately protect 
Trifolium friscanum from the impacts of mining on private lands. The 
majority of individuals (3 populations) have the potential to be 
impacted by mining in the future. However, because none of the 
populations are directly impacted by current mining levels on SITLA or 
private lands, we consider threats under Factor D to be moderate at 
this time.
    We consider the magnitude of Factor E moderate, because, although 
small population size and climate change make the species intrinsically 
more vulnerable, we are uncertain of whether they would rise to the 
level of threat by themselves. However, when collectively analyzed with 
the threats listed under Factor A, they may rise to the level of threat 
in the foreseeable future. Although we are uncertain about the direct 
impacts of global climate change on Trifolium friscanum, we expect the 
species to respond negatively to changed environmental conditions and 
drought, especially when combined with the effects of small population 
size and the threat of increased mining activities.
    Therefore, we consider the threats that Trifolium friscanum faces 
to be moderate in magnitude because the major threats (mining, 
nonnative invasive species, small population size, plus inadequacy of 
existing regulatory mechanisms), while serious and occurring rangewide, 
do not collectively rise to the level of high magnitude.
    Under our LPN guidance, the second criterion we consider in 
assigning a listing priority is the immediacy of threats. This 
criterion is intended to ensure that the species facing actual, 
identifiable threats are given priority over those for which threats 
are only potential or those that are intrinsically vulnerable but are 
not known to be presently facing such threats. We consider all of the 
threats to be imminent because we have factual information that the 
threats are identifiable and that the species is currently facing them 
in many portions of its range. These actual, identifiable threats are 
covered in greater detail in Factors A, D, and E of this finding. The 
majority of threats are ongoing and, therefore, imminent, although 
mining is currently ongoing in the habitat of only one of the 
populations. In addition to their current existence, we expect these 
threats, except for inadequate regulations, to continue and likely 
intensify in the foreseeable future.
    The third criterion in our LPN guidance is intended to devote 
resources to those species representing highly distinctive or isolated 
gene pools as reflected by taxonomy. Trifolium friscanum is a valid 
taxon at the species level and, therefore, receives a higher priority 
than subspecies, but a lower priority than species in a monotypic 
genus. Therefore, we assigned T. friscanum an LPN of 8.
    We will continue to monitor the threats to Trifolium friscanum and 
the species' status on an annual basis, and, should the magnitude or 
the imminence of the threats change, we will revisit our assessment of 
the LPN.
    While we conclude that listing Trifolium friscanum is warranted, an 
immediate proposal to list this species is precluded by other higher 
priority listings, which we address in the Preclusion and Expeditious 
Progress section below. Because we have assigned T. friscanum an LPN of 
8, work on a proposed listing determination for T. friscanum is 
precluded by work on higher priority listing actions with absolute 
statutory, court-ordered, or court-approved deadlines and final listing 
determinations for those species that were proposed for listing with

[[Page 10198]]

funds from FY 2010. This work includes all the actions listed in the 
tables below under expeditious progress.

Preclusion and Expeditious Progress

    Preclusion is a function of the listing priority of a species in 
relation to the resources that are available and competing demands for 
those resources. Thus, in any given fiscal year, multiple factors 
dictate whether it will be possible to undertake work on a proposed 
listing regulation or whether promulgation of such a proposal is 
warranted but precluded by higher priority listing actions.
    The resources available for listing actions are determined through 
the annual Congressional appropriations process. The appropriation for 
the Services' Listing Program is available to support work involving 
the following listing actions: Proposed and final listing rules; 90-day 
and 12-month findings on petitions to add species to the Lists of 
Endangered and Threatened Wildlife and Plants (Lists) or to change the 
status of a species from threatened to endangered; annual 
determinations on prior ``warranted but precluded'' petition findings 
as required under section 4(b)(3)(C)(i) of the ESA; critical habitat 
petition findings; proposed and final rules designating critical 
habitat; and litigation-related, administrative, and program-management 
functions (including preparing and allocating budgets, responding to 
Congressional and public inquiries, and conducting public outreach 
regarding listing and critical habitat).
    The work involved in preparing various listing documents can be 
extensive and may include, but is not limited to: Gathering and 
assessing the best scientific and commercial data available and 
conducting analyses used as the basis for our decisions; writing and 
publishing documents; and obtaining, reviewing, and evaluating public 
comments and peer review comments on proposed rules and incorporating 
relevant information into final rules. The number of listing actions 
that we can undertake in a given year also is influenced by the 
complexity of those listing actions; that is, more complex actions 
generally are more costly. For example, during the past several years 
the cost (excluding publication costs) for preparing a 12-month 
finding, without a proposed rule, has ranged from approximately $11,000 
for one species with a restricted range and involving a relatively 
uncomplicated analysis to $305,000 for another species that is wide 
ranging and involving a complex analysis.
    We cannot spend more than is appropriated for the Listing Program 
without violating the Anti-Deficiency Act (see 31 U.S.C. 
1341(a)(1)(A)). In addition, in FY 1998 and for each FY since then, 
Congress has placed a statutory cap on funds which may be expended for 
the Listing Program, equal to the amount expressly appropriated for 
that purpose in that FY. This cap was designed to prevent funds 
appropriated for other functions under the ESA (for example, recovery 
funds for removing species from the Lists), or for other Service 
programs, from being used for Listing Program actions (see House Report 
105-163, 105th Congress, 1st Session, July 1, 1997).
    Recognizing that designation of critical habitat for species 
already listed would consume most of the overall Listing Program 
appropriation, Congress also put a critical habitat subcap in place in 
FY 2002 and has retained it each subsequent year to ensure that some 
funds are available for other work in the Listing Program: ``The 
critical habitat designation subcap will ensure that some funding is 
available to address other listing activities'' (House Report No. 107-
103, 107th Congress, 1st Session, June 19, 2001). In FY 2002 and each 
year until FY 2006, the Service has had to use virtually the entire 
critical habitat subcap to address court-mandated designations of 
critical habitat, and consequently none of the critical habitat subcap 
funds have been available for other listing activities. In FY 2007, we 
were able to use some of the critical habitat subcap funds to fund 
proposed listing determinations for high-priority candidate species. In 
FY 2009, while we were unable to use any of the critical habitat subcap 
funds to fund proposed listing determinations, we did use some of this 
money to fund the critical habitat portion of some proposed listing 
determinations so that the proposed listing determination and proposed 
critical habitat designation could be combined into one rule, thereby 
being more efficient in our work. In FY 2010, we are using some of the 
critical habitat subcap funds to fund actions with statutory deadlines.
    Thus, through the listing cap, the critical habitat subcap, and the 
amount of funds needed to address court-mandated critical habitat 
designations, Congress and the courts have in effect determined the 
amount of money available for other listing activities. Therefore, the 
funds in the listing cap, other than those needed to address court-
mandated critical habitat for already listed species, set the limits on 
our petition finding determinations.
    Congress also recognized that the availability of resources was the 
key element in deciding, when making a 12-month petition finding, 
whether we would prepare and issue a listing proposal or instead make a 
``warranted but precluded'' finding for a given species. The Conference 
Report accompanying Public Law 97-304, which established the current 
statutory deadlines and the warranted-but-precluded finding, states (in 
a discussion on 90-day petition findings that by its own terms also 
covers 12-month findings) that the deadlines were ``not intended to 
allow the Secretary to delay commencing the rulemaking process for any 
reason other than that the existence of pending or imminent proposals 
to list species subject to a greater degree of threat would make 
allocation of resources to such a petition [that is, for a lower-
ranking species] unwise.''
    In FY 2010, expeditious progress is that amount of work that can be 
achieved with $10,471,000, which is the amount of money that Congress 
appropriated for the Listing Program (that is, the portion of the 
Listing Program funding not related to critical habitat designations 
for species that are already listed). However, these funds are not 
enough to fully fund all our court-ordered and statutory listing 
actions in FY 2010, so we are using $1,114,417 of our critical habitat 
subcap funds in order to work on all of our required petition findings 
and listing determinations. This brings the total amount of funds we 
have for listing actions in FY 2010 to $11,585,417.
    Starting in FY 2010, we also are using our funds to work on listing 
actions for foreign species, because that work was transferred from the 
Division of Scientific Authority, International Affairs Program, to the 
Endangered Species Program. Our process is to make our determinations 
of preclusion on a nationwide basis to ensure that the species most in 
need of listing will be addressed first and also because we allocate 
our listing budget on a nationwide basis. The $11,585,417 is being used 
to fund work in the following categories: Compliance with court orders 
and court-approved settlement agreements requiring that petition 
findings or listing determinations be completed by a specific date; 
section 4 (of the ESA) listing actions with absolute statutory 
deadlines; essential litigation-related, administrative, and listing 
program-management functions; and high-priority listing actions for 
some of our candidate species. The allocations for each specific 
listing action are identified

[[Page 10199]]

in the Service's FY 2010 Allocation Table (part of our administrative 
record).
    In FY 2007, we had more than 120 species with an LPN of 2, based on 
our September 21, 1983, guidance for assigning an LPN for each 
candidate species (48 FR 43098). Using this guidance, we assign each 
candidate an LPN of 1 to 12, depending on the magnitude of threats 
(high vs. moderate to low), immediacy of threats (imminent or 
nonimminent), and taxonomic status of the species (in order of 
priority: monotypic genus (a species that is the sole member of a 
genus); species; or part of a species (subspecies, DPS, or significant 
portion of the range)). The lower the listing priority number, the 
higher the listing priority (that is, a species with an LPN of 1 would 
have the highest listing priority). Because of the large number of 
high-priority species, we further ranked the candidate species with an 
LPN of 2 by using the following extinction-risk type criteria: 
International Union for the Conservation of Nature and Natural 
Resources (IUCN) Red list status/rank, Heritage rank (provided by 
NatureServe), Heritage threat rank (provided by NatureServe), and 
species currently with fewer than 50 individuals, or 4 or fewer 
populations. Those species with the highest IUCN rank (critically 
endangered), the highest Heritage rank (G1), the highest Heritage 
threat rank (substantial, imminent threats), and currently with fewer 
than 50 individuals, or fewer than 4 populations, comprised a group of 
approximately 40 candidate species (``Top 40''). These 40 candidate 
species have had the highest priority to receive funding to work on a 
proposed listing determination. As we work on proposed and final 
listing rules for these 40 candidates, we are applying the ranking 
criteria to the next group of candidates with an LPN of 2 and 3 to 
determine the next set of highest priority candidate species.
    To be more efficient in our listing process, as we work on proposed 
rules for these species in the next several years, we are preparing 
multi-species proposals when appropriate, and these may include species 
with lower priority if they overlap geographically or have the same 
threats as a species with an LPN of 2. In addition, available staff 
resources also are a factor in determining high-priority species 
provided with funding. Finally, proposed rules for reclassification of 
threatened species to endangered are lower priority, since as listed 
species, they are already afforded the protection of the ESA and 
implementing regulations.
    We assigned Eriogonum soredium, Lepidium ostleri and Trifolium 
friscanum an LPN of 8. This is based on our finding that the species 
face immediate and moderate magnitude threats from the present or 
threatened destruction, modification or curtailment of its habitat; the 
inadequacy of existing regulatory mechanisms; and other natural or man-
made factors affecting their continued existence. These threats are 
ongoing and, in some cases (e.g., nonnative species), considered 
irreversible. Under our 1983 Guidelines, a ``species'' facing imminent 
moderate-magnitude threats is assigned an LPN of 7, 8, or 9 depending 
on its taxonomic status. Because E. soredium, L. ostleri and T. 
friscanum are species, we assigned an LPN of 8 to each. Therefore, work 
on a proposed listing determination for E. soredium, L. ostleri and T. 
friscanum is precluded by work on higher priority candidate species 
(i.e., species with LPN of 7); listing actions with absolute statutory, 
court ordered, or court-approved deadlines; and final listing 
determinations for those species that were proposed for listing with 
funds from previous FYs. This work includes all the actions listed in 
the tables below under expeditious progress.
    As explained above, a determination that listing is warranted but 
precluded also must demonstrate that expeditious progress is being made 
to add or remove qualified species to and from the Lists of Endangered 
and Threatened Wildlife and Plants. (Although we do not discuss it in 
detail here, we also are making expeditious progress in removing 
species from the Lists under the Recovery program, which is funded by a 
separate line item in the budget of the Endangered Species Program. As 
explained above in our description of the statutory cap on Listing 
Program funds, the Recovery Program funds and actions supported by them 
cannot be considered in determining expeditious progress made in the 
Listing Program.) As with our ``precluded'' finding, expeditious 
progress in adding qualified species to the Lists is a function of the 
resources available and the competing demands for those funds. Given 
that limitation, we find that we are making progress in FY 2010 in the 
Listing Program. This progress included preparing and publishing the 
following determinations:

                                        FY 2010 Completed Listing Actions
----------------------------------------------------------------------------------------------------------------
      Publication date                  Title                   Actions              Federal Register pages
----------------------------------------------------------------------------------------------------------------
10/08/2009.................  Listing Lepidium            Final Listing,         74 FR 52013-52064.
                              papilliferum (Slickspot     Threatened.
                              Peppergrass) as a
                              Threatened Species
                              Throughout Its Range.
10/27/2009.................  90-day Finding on a         Notice of 90-day       74 FR 55177-55180.
                              Petition To List the        Petition Finding,
                              American Dipper in the      Not substantial.
                              Black Hills of SD as
                              Threatened or Endangered.
10/28/2009.................  Status Review of Arctic     Notice of Intent to    74 FR 55524-55525.
                              Grayling (Thymallus         Conduct Status
                              arcticus) in the Upper      Review.
                              Missouri River System.
11/03/2009.................  Listing the British         Proposed Listing       74 FR 56757-56770.
                              Columbia DPS of the Queen   Threatened.
                              Charlotte Goshawk Under
                              the ESA: Proposed rule.
11/03/2009.................  Listing the Salmon-Crested  Proposed Listing       74 FR 56770-56791.
                              Cockatoo as Threatened      Threatened.
                              Throughout Its Range with
                              Special Rule.
11/23/2009.................  Status Review of Gunnison   Notice of Intent to    74 FR 61100-61102.
                              sage-grouse (Centrocercus   Conduct Status
                              minimus).                   Review.
12/03/2009.................  12-Month Finding on a       Notice of 12-month     74 FR 63343-63366.
                              Petition to List the        petition finding,
                              Black-tailed Prairie Dog    Not warranted.
                              as Threatened or
                              Endangered.
12/03/2009.................  90-Day Finding on a         Notice of 90-day       74 FR 63337-63343.
                              Petition to List            Petition Finding,
                              Sprague's Pipit as          Substantial.
                              Threatened or Endangered.
12/15/2009.................  90-Day Finding on           Notice of 90-day       74 FR 66260-66271.
                              Petitions To List Nine      Petition Finding,
                              Species of Mussels From     Substantial.
                              TX as Threatened or
                              Endangered With Critical
                              Habitat.
12/16/2009.................  Partial 90-Day Finding on   Notice of 90-day       74 FR 66865-66905.
                              a Petition to List 475      Petition Finding,
                              Species in the              Not substantial &
                              Southwestern U.S. as        Substantial.
                              Threatened or Endangered
                              With Critical Habitat.

[[Page 10200]]

 
12/17/2009.................  12-month Finding on a       Notice of 12-month     74 FR 66937-66950.
                              Petition To Change the      petition finding,
                              Final Listing of the DPS    Warranted but
                              of the Canada Lynx To       precluded.
                              Include NM.
01/05/2010.................  Listing Foreign Bird        Proposed Listing,      75 FR 605-649.
                              Species in Peru & Bolivia   Endangered.
                              as Endangered Throughout
                              Their Range.
01/05/2010.................  Listing Six Foreign Birds   Proposed Listing,      75 FR 286-310.
                              as Endangered Throughout    Endangered.
                              Their Range.
01/05/2010.................  Withdrawal of Proposed      Proposed rule,         75 FR 310-316.
                              Rule to List Cook's         withdrawal.
                              Petrel.
01/05/2010.................  Final Rule to List the      Final Listing,         75 FR 235-250.
                              Galapagos Petrel &          Threatened.
                              Heinroth's Shearwater as
                              Threatened Throughout
                              Their Ranges.
01/20/2010.................  Initiation of Status        Notice of Intent to    75 FR 3190-3191.
                              Review for Agave            Conduct Status
                              eggersiana & Solanum        Review.
                              conocarpum.
02/09/2010.................  12-month Finding on a       Notice of 12-month     75 FR 6437-6471.
                              Petition to List the        petition finding,
                              American Pika as            Not warranted.
                              Threatened or Endangered.
02/25/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 8601-8621.
                              Petition To List the        petition finding,
                              Sonoran Desert Population   Not warranted.
                              of the Bald Eagle as a
                              Threatened or Endangered
                              DPS.
02/25/2010.................  Withdrawal of Proposed      Withdrawal of          75 FR 8621-8644.
                              Rule To List the            Proposed Rule to
                              Southwestern Washington/    List.
                              Columbia River DPS of
                              Coastal Cutthroat Trout
                              (Oncorhynchus clarki
                              clarki) as Threatened.
03/18/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 13068-13071.
                              Petition to List the        Petition Finding,
                              Berry Cave Salamander as    Substantial.
                              Endangered.
03/23/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 13717-13720.
                              Petition to List the        Petition Finding,
                              Southern Hickorynut         Not substantial.
                              Mussel (Obovaria
                              jacksoniana) as
                              Endangered or Threatened.
03/23/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 13720-13726.
                              Petition to List the        Petition Finding,
                              Striped Newt as             Substantial.
                              Threatened.
03/23/2010.................  12-Month Findings for       Notice of 12-month     75 FR 13910-14014.
                              Petitions to List the       petition finding,
                              Greater Sage-Grouse         Warranted but
                              (Centrocercus               precluded.
                              urophasianus) as
                              Threatened or Endangered.
03/31/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 16050-16065.
                              Petition to List the        petition finding,
                              Tucson Shovel-Nosed Snake   Warranted but
                              (Chionactis occipitalis     precluded.
                              klauberi) as Threatened
                              or Endangered with
                              Critical Habitat.
04/05/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 17062-17070.
                              Petition To List Thorne's   Petition Finding,
                              Hairstreak Butterfly as     Substantial.
                              or Endangered.
04/06/2010.................  12-month Finding on a       Notice of 12-month     75 FR 17352-17363.
                              Petition To List the        petition finding,
                              Mountain Whitefish in the   Not warranted.
                              Big Lost River, ID, as
                              Endangered or Threatened.
04/06/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 17363-17367.
                              Petition to List a          Petition Finding,
                              Stonefly (Isoperla          Not substantial.
                              jewetti) and a Mayfly
                              (Fallceon eatoni) as
                              Threatened or Endangered
                              with Critical Habitat.
04/07/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 17667-17680.
                              Petition to Reclassify      petition finding,
                              the Delta Smelt From        Warranted but
                              Threatened to Endangered    precluded.
                              Throughout Its Range.
04/13/2010.................  Determination of            Final Listing,         75 FR 18959-19165.
                              Endangered Status for 48    Endangered.
                              Species on Kauai &
                              Designation of Critical
                              Habitat.
04/15/2010.................  Initiation of Status        Notice of Initiation   75 FR 19591-19592.
                              Review of the North         of Status Review.
                              American Wolverine in the
                              Contiguous U.S.
04/15/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 19592-19607.
                              Petition to List the        petition finding,
                              Wyoming Pocket Gopher as    Not warranted.
                              Endangered or Threatened
                              with Critical Habitat.
04/16/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 19925-19935.
                              Petition to List a DPS of   Petition Finding,
                              the Fisher in Its U.S.      Substantial.
                              Northern Rocky Mountain
                              Range as Endangered or
                              Threatened with Critical
                              Habitat.
04/20/2010.................  Initiation of Status        Notice of Initiation   75 FR 20547-20548.
                              Review for Sacramento       of Status Review.
                              splittail (Pogonichthys
                              macrolepidotus).
04/26/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 21568-21571.
                              Petition to List the        Petition Finding,
                              Harlequin Butterfly as      Substantial.
                              Endangered.
04/27/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 22012-22025.
                              Petition to List Susan's    petition finding,
                              Purse-making Caddisfly      Not warranted.
                              (Ochrotrichia susanae) as
                              Threatened or Endangered.
04/27/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 22063-22070.
                              Petition to List the        Petition Finding,
                              Mohave Ground Squirrel as   Substantial.
                              Endangered with Critical
                              Habitat.
05/04/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 23654-23663.
                              Petition to List Hermes     Petition Finding,
                              Copper Butterfly as         Substantial.
                              Threatened or Endangered.
06/01/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 30313-30318.
                              Petition To List Castanea   Petition Finding,
                              pumila var. ozarkensis.     Substantial.
06/01/2010.................  12-Month Finding on a       Notice of 12-month     75 FR 30338-30363.
                              Petition to List the        petition finding,
                              White-tailed Prairie Dog    Not warranted.
                              as Endangered or
                              Threatened.
06/09/2010.................  90-Day Finding on a         Notice of 90-day       75 FR 32728-32734.
                              Petition To List van        Petition Finding,
                              Rossem's Gull-billed Tern   Substantial.
                              as Endangered or
                              Threatened.
----------------------------------------------------------------------------------------------------------------

    Our expeditious progress also includes work on listing actions that 
we funded in FY 2010 but have not yet been completed to date. These 
actions are listed below. Actions in the top section of the table are 
being conducted under a deadline set by a court. Actions in the middle 
section of the table are being conducted to meet statutory timelines, 
that is, timelines required under the ESA. Actions in the bottom 
section of the table are high-priority listing actions. These actions 
include work primarily on species with an LPN of 2, and selection of 
these species is

[[Page 10201]]

partially based on available staff resources, and when appropriate, 
include species with a lower priority if they overlap geographically or 
have the same threats as the species with the high priority. Including 
these species together in the same proposed rule results in 
considerable savings in time and funding, as compared to preparing 
separate proposed rules for each of them in the future.

                                 Actions Funded in FY 2010 but Not Yet Completed
----------------------------------------------------------------------------------------------------------------
                    Species                                                  Action
----------------------------------------------------------------------------------------------------------------
Actions Subject to Court Order/Settlement
 Agreement:
    6 Birds from Eurasia......................  Final listing determination.
    Flat-tailed horned lizard.................  Final listing determination.
    Mountain plover...........................  Final listing determination.
    6 Birds from Peru.........................  Proposed listing determination.
    Sacramento splittail......................  Proposed listing determination.
    Gunnison sage-grouse......................  12-month petition finding.
    Wolverine.................................  12-month petition finding.
    Montana Arctic grayling...................  12-month petition finding.
    Agave eggersiana..........................  12-month petition finding.
    Solanum conocarpum........................  12-month petition finding.
    Mountain plover...........................  12-month petition finding.
    Thorne's Hairstreak Butterfly.............  12-month petition finding.
    Hermes copper butterfly...................  12-month petition finding.
Actions With Statutory Deadlines:
    Casey's june beetle.......................  Final listing determination.
    Georgia pigtoe, interrupted rocksnail, and  Final listing determination.
     rough hornsnail.
    2 Hawaiian damselflies....................  Final listing determination.
    African penguin...........................  Final listing determination.
    3 Foreign bird species (Andean flamingo,    Final listing determination.
     Chilean woodstar, St. Lucia forest
     thrush).
    5 Penguin species.........................  Final listing determination.
    Southern rockhopper penguin--Campbell       Final listing determination.
     Plateau population.
    5 Bird species from Colombia and Ecuador..  Final listing determination.
    7 Bird species from Brazil................  Final listing determination.
    Queen Charlotte goshawk...................  Final listing determination.
    Salmon crested cockatoo...................  Proposed listing determination.
    Black-footed albatross....................  12-month petition finding.
    Mount Charleston blue butterfly...........  12-month petition finding.
    Least chub \1\............................  12-month petition finding.
    Mojave fringe-toed lizard \1\.............  12-month petition finding.
    Pygmy rabbit (rangewide) \1\..............  12-month petition finding.
    Kokanee--Lake Sammamish population \1\....  12-month petition finding.
    Delta smelt (uplisting)...................  12-month petition finding.
    Cactus ferruginous pygmy-owl \1\..........  12-month petition finding.
    Northern leopard frog.....................  12-month petition finding.
    Tehachapi slender salamander..............  12-month petition finding.
    Coqui Llanero.............................  12-month petition finding.
    White-sided jackrabbit....................  12-month petition finding.
    Jemez Mountains salamander................  12-month petition finding.
    Dusky tree vole...........................  12-month petition finding.
    Eagle Lake trout \1\......................  12-month petition finding.
    29 of 206 species.........................  12-month petition finding.
    Desert tortoise--Sonoran population.......  12-month petition finding.
    Gopher tortoise--eastern population.......  12-month petition finding.
    Amargosa toad.............................  12-month petition finding.
    Pacific walrus............................  12-month petition finding.
    Wrights marsh thistle.....................  12-month petition finding.
    67 of 475 southwest species...............  12-month petition finding.
    9 Southwest mussel species................  12-month petition finding.
    14 parrots (foreign species)..............  12-month petition finding.
    Berry Cave salamander \1\.................  12-month petition finding.
    Striped Newt \1\..........................  12-month petition finding.
    Fisher--Northern Rocky Mountain Range \1\.  12-month petition finding.
    Mohave Ground Squirrel \1\................  12-month petition finding.
    Puerto Rico Harlequin Butterfly...........  12-month petition finding.
    Western gull-billed tern..................  12-month petition finding.
    Ozark chinquapin (Castanea pumila var.      12-month petition finding.
     ozarkensis).
    Southeastern population of snowy plover     90-day petition finding.
     and wintering population of piping plover
     \1\.
    Eagle Lake trout \1\......................  90-day petition finding.
    Smooth-billed ani \1\.....................  90-day petition finding.
    Bay Springs salamander \1\................  90-day petition finding.
    32 species of snails and slugs \1\........  90-day petition finding.
    Calopogon oklahomensis \1\................  90-day petition finding.
    White-bark pine...........................  90-day petition finding.
    42 snail species (Nevada and Utah)........  90-day petition finding.
    HI yellow-faced bees......................  90-day petition finding.
    Red knot roselaari subspecies.............  90-day petition finding.

[[Page 10202]]

 
    Honduran emerald..........................  90-day petition finding.
    Peary caribou.............................  90-day petition finding.
    Plains bison..............................  90-day petition finding.
    Giant Palouse earthworm...................  90-day petition finding.
    Mexican gray wolf.........................  90-day petition finding.
    Spring Mountains checkerspot butterfly....  90-day petition finding.
    Spring pygmy sunfish......................  90-day petition finding.
    San Francisco manzanita...................  90-day petition finding.
    Bay skipper...............................  90-day petition finding.
    Unsilvered fritillary.....................  90-day petition finding.
    Texas kangaroo rat........................  90-day petition finding.
    Spot-tailed earless lizard................  90-day petition finding.
    Eastern small-footed bat..................  90-day petition finding.
    Northern long-eared bat...................  90-day petition finding.
    Prairie chub..............................  90-day petition finding.
    10 species of Great Basin butterfly.......  90-day petition finding.
    6 sand dune (scarab) beetles..............  90-day petition finding.
    Golden-winged warbler.....................  90-day petition finding.
    Sand-verbena moth.........................  90-day petition finding.
    Aztec (beautiful) gilia...................  90-day petition finding.
    Arapahoe snowfly..........................  90-day petition finding.
High-Priority Listing Actions: \3\
    19 Oahu candidate species \3\ (16 plants,   Proposed listing.
     3 damselflies) (15 with LPN = 2, 3 with
     LPN = 3, 1 with LPN = 9).
    17 Maui-Nui candidate species \3\ (14       Proposed listing.
     plants, 3 tree snails) (12 with LPN = 2,
     2 with LPN = 3, 3 with LPN = 8).
    Sand dune lizard \3\ (LPN = 2)............  Proposed listing.
    2 Arizona springsnails \3\ (Pyrgulopsis     Proposed listing.
     bernadina (LPN = 2), Pyrgulopsis
     trivialis (LPN = 2).
    2 New Mexico springsnails \3\ (Pyrgulopsis  Proposed listing.
     chupaderae (LPN = 2), Pyrgulopsis
     thermalis (LPN = 11).
    2 mussels \3\ (rayed bean (LPN = 2),        Proposed listing.
     snuffbox (No LPN).
    2 mussels \3\ (sheepnose (LPN = 2),         Proposed listing.
     spectaclecase (LPN = 4)).
    Ozark hellbender \2\ (LPN = 3)............  Proposed listing.
    Altamaha spinymussel \3\ (LPN = 2)........  Proposed listing.
    5 southeast fish \3\ (rush darter (LPN =    Proposed listing.
     2), chucky madtom (LPN = 2), yellowcheek
     darter (LPN = 2), Cumberland darter (LPN
     = 5), laurel dace (LPN = 5).
    8 southeast mussels (southern kidneyshell   Proposed listing.
     (LPN = 2), round ebonyshell (LPN = 2),
     Alabama pearlshell (LPN = 2), southern
     sandshell (LPN = 5), fuzzy pigtoe (LPN =
     5), Choctaw bean (LPN = 5), narrow pigtoe
     (LPN = 5), & tapered pigtoe (LPN = 11)).
    3 Colorado plants \3\ (Pagosa skyrocket     Proposed listing.
     (Ipomopsis polyantha) (LPN = 2), Parchute
     beardtongue (Penstemon debilis) (LPN =
     2), Debeque phacelia (Phacelia submutica)
     (LPN = 8)).
    2 Texas plants (Texas golden gladecress     Proposed listing.
     (Leavenworthia texana) (LPN = 2), Neches
     River rose mallow (Hibiscus dasycalyx)
     (LPN = 5)).
    Florida bonneted bat (LPN = 2)............  Proposed listing.
    Kittlitz's murrelet (LPN = 2).............  Proposed listing.
----------------------------------------------------------------------------------------------------------------
\1\ Funds for listing actions for these species were provided in previous FYs.
\2\ We funded a proposed rule for this subspecies with an LPN of 3 ahead of other species with LPN of 2, because
  the threats to the species were so imminent and of a high magnitude that we considered emergency listing if we
  were unable to fund work on a proposed listing rule in FY 2008.
\3\ Funds for these high-priority listing actions were provided in FY 2008 or 2009.

    We have endeavored to make our listing actions as efficient and 
timely as possible, given the requirements of the relevant law and 
regulations, and constraints relating to workload and personnel. We are 
continually considering ways to streamline processes or achieve 
economies of scale, such as by batching related actions together. Given 
our limited budget for implementing section 4 of the ESA, these actions 
described above collectively constitute expeditious progress.
    Eriogonum soredium, Lepidium ostleri, and Trifolium friscanum will 
be added to the list of candidate species upon publication of this 12-
month finding. We will continue to monitor the status of these species 
as new information becomes available. This review will determine if a 
change in status is warranted, including the need to make prompt use of 
emergency listing procedures.
    We intend that any proposed listing action for Eriogonum soredium, 
Lepidium ostleri, and Trifolium friscanum will be as accurate as 
possible. Therefore, we will continue to accept additional information 
and comments from all concerned governmental agencies, the scientific 
community, industry, or any other interested party concerning this 
finding.

References Cited

    A complete list of references cited is available on the Internet at 
http://www.regulations.gov or upon request from the Utah Ecological 
Services Field Office (see ADDRESSES section).

Authors

    The primary authors of this notice are the staff members of the 
Utah Ecological Services Field Office.

Authority

    The authority for this section is section 4 of the Endangered 
Species Act of 1973, as amended (16 U.S.C. 1531 et seq.).


[[Page 10203]]


    Dated: February 2, 2011.
Rowan W. Gould,
Acting Director, Fish and Wildlife Service.
[FR Doc. 2011-3675 Filed 2-22-11; 8:45 am]
BILLING CODE 4310-55-P