<<

For U.S. Fish and Wildlife Service Section 7 Consultation

FORMAL CONSULTATION FY 2013

Lower Steamboat Timber Sale and Lonely Oak Project

UMPQUA NATIONAL FOREST

Prepared by:_/s/Josh Chapman______Date:_9/20/2013______Josh Chapman Level One Team Wildlife Biologist Umpqua National Forest

Table of Contents I. Introduction ...... 4 II. Status Of The Northern Spotted ...... 5 III. Description of the Projects Being Analyzed for this Biological Assessment ..... 39 A. Lower Steamboat Timber Sale ...... 39 i. Lower Steamboat Timber Sale Action Area ...... 39 ii. Status of the in the Lower Steamboat Action Area ...... 42 iii. Lower Steamboat Timber Sale Proposed Action Description ...... 43 B. Lonely Oak Project ...... 46 i. Lonely Oak Project Action Area ...... 46 ii. Status of the Species in the Lonely Oak Action Area ...... 49 iii. Lonely Oak Project Proposed Action Description ...... 50 IV. Direct And Indirect Effects ...... 51 A. Lower Steamboat ...... 51 B. Lonely Oak ...... 56 V. Direct and Indirect Effects to Critical Habitat ...... 59 VI. Consistency with the Recovery Plan ...... 62 VIII. Cumulative Effects ...... 62 IX. Interrelated and Interdependent Actions ...... 63 X. Determinations ...... 63 XI. Literature Cited ...... 64 Appendix A. Project Design Features for Parker Wyatt Timber Sale ...... 74

Tables

Table 1. Range-wide Aggregate of Changes to NRF1 Habitat Acres from Activities Subject to Section 7 Consultations and Other Causes...... 25 Table 2. Aggregate Results of All Adjusted, Suitable Habitat (NRF1) Acres Affected by Section 7 Consultation on NWFP Lands for the ; Baseline and Summary of Effects by State, Physiographic Province and Land Use Function...... 26 Table 3. Demographic results from spotted owl demographic study areas (adapted from Forsman et al. 2011)...... 30 Table 4. Environmental Baseline for Revised Northern Spotted Owl Critical Habitat for the West Cascades South Unit, Sub-Unit WCS-5...... 35 Table 5. Environmental Baseline: Umpqua National Forest environmental baseline as of March 2013, utilizing the 15 year monitoring report for the spotted owl habitat model (Davis 2011)...... 37 Table 6. Description of the total Action Area environmental baseline for the Lower Steamboat Timber Sale by land use allocation and amount of suitable and capable habitat present utilizing 1 the 15 year monitoring report for the Northwest Forest Plan spotted owl habitat model (Davis 2011)...... 41 Table 7. Northern Spotted Owl activity centers within the Lower Steamboat Action Area listed by ID number, with percent of nesting, roosting and foraging (NRF) habitat at the 800m core scale and 1.2 mile home range scale...... 43 Table 8. Acres of proposed treatment by fuels and silvicultural prescription with 2012 Northern Spotted Owl Critical Habitat...... 45 Table 9. Silviculture and fuels treatment acres, pre and post treatment canopy covers, and amounts and sizes of gaps proposed as part of the Lower Steamboat Timber Sale...... 46 Table 10. Description of the total Action Area environmental baseline for the Lonely Oak Timber Sale by land use allocation and amount of suitable and capable habitat present utilizing the 15 year monitoring report for the Northwest Forest Plan spotted owl habitat model (Davis 2011)...... 48 Table 11. Northern Spotted Owl activity centers within the Lower Steamboat Action Area listed by ID number, with percent of nesting, roosting and foraging (NRF) habitat at the 800m core scale and 1.2 mile home range scale...... 49 Table 12. Acres of treatment proposed as part of the Lonely Oak Project...... 50 Table 13. Acres by treatment type within 2012 Northern Spotted Owl Critical Habitat that are part of the Lonely Oak Project...... 50 Table 14. Table of impacts to dispersal habitat resulting from the Lower Steamboat Timber Sale...... 52 Table 15. Table of all owl sites amount of dispersal habitat pre and post treatment within the Lower Steamboat Timber Sale action area...... 52 Table 16. Disruption distances for the northern spotted owl during the breeding period of March 1-September 30...... 54 Table 17. Amount of NRF and Dispersal habitat maintained and amount of NRF habitat downgraded to dispersal habitat as a result of the Lonely Oak project...... 56 Table 18. Amounts of nesting, roosting and foraging habitat at the home range (1.2 mile), core (800m) and nest patch (70m) scale pre and post Lonely Oak Project treatments for the KPAC’s within the Lonely Oak Project Action area...... 56 Table 19. Total impacts to Northern Spotted Owl dispersal and nesting, roosting and foraging habitat for both the Lower Steamboat Timber Sale and the Lonely Oak Project...... 63

Figures Figure 1. Locator map for the Lower Steamboat Timber Sale and Lonely Oak Project, including the adjacent National Forests and Final 2012 Critical Habitat Units...... 4 Figure 2. The 12 physiographic Provinces within the range of the northern spotted owl (USDI FWS 1992a, p. 31)...... 7 Figure 3. Northern spotted owl demographic study areas (Forsman et al. 2011)...... 29 Figure 4. Spotted owl territories and reproduction and territories relationship on the Tyee Density Study Area (Forsman et al. 2013)...... 33

2

Figure 5. Percentage of sites occupied by pairs of northern spotted compared among land use allocations in the HJ Andrews-Central Cascades study area, Willamette National Forest, from 1997 – 2012 (Dugger et al. 2013)...... 34 Figure 6. Map of the Northwest Forest Plan Land Use Allocations on and adjacent to the Umpqua National Forest and the Lower Steamboat Timber Sale and Lonely Oak Project, as well as the 2012 Critical Habitat Sub-Unit boundaries...... 38 Figure 7. Map of the Lower Steamboat Timber Sale action area and harvest units, along with 2012 Northern Spotted Owl Critical Habitat and owl cores...... 40 Figure 8. Lower Steamboat timber sale action area and all associated owl sites within the action area, along with the Nesting Roosting and Foraging (NRF) habitat layer...... 42 Figure 9. Lonely Oak Project action area and treatment units, along with 2012 Northern Spotted Owl Critical Habitat and 100 acre late successional reserve owl cores...... 47 Figure 10. Lonely Oak project action area and all associated owl sites within the action area, along with the Nesting Roosting and Foraging (NRF) habitat layer...... 49 Figure 11. Lower Steamboat Timber Sale disturbance map showing the 60m and quarter mile disturbance buffers around all KPAC’s within the action area...... 55 Figure 12. Lonely Oak project disturbance map showing the 60m and quarter mile disturbance buffers around all KPAC’s within the action area...... 58

3

I. Introduction

Under the Endangered Species Act of 1973 (as amended), Federal agencies are required to review all actions authorized, funded or carried out by them to ensure that those actions do not jeopardize the continued existence of any Federally listed species (Section 7(a)(2)). The purpose of this Biological Assessment (BA) is to meet the Umpqua National Forest’s requirements to consult with the U.S. Fish and Wildlife Service and disclose all potential direct, indirect and cumulative effects to the Federally Threatened Northern Spotted Owl ( occidentalis caurina) as a result of the Lower Steamboat and Lonely Oak Timber Sales (Figure 1). This BA has been reviewed by the U.S. Fish and Wildlife Service and Umpqua National Forest Level One consultation members in preparation for formal submittal.

Figure 1. Locator map for the Lower Steamboat Timber Sale and Lonely Oak Project, including the adjacent National Forests and Final 2012 Critical Habitat Units.

4

II. Status Of The Northern Spotted Owl

Legal Status

The spotted owl was listed as threatened on June 26, 1990 due to widespread loss and adverse modification of suitable habitat across the owl’s entire range and the inadequacy of existing regulatory mechanisms to conserve the owl (USDI FWS 1990, p. 26114). The U.S. Fish and Wildlife Service recovery priority number for the spotted owl is 12C (USDI FWS 2011a, p. I6), on a scale of 1C (highest) to 18 (lowest). This number reflects a moderate degree of threat, a low potential for recovery, the spotted owl’s taxonomic status as a subspecies and inherent conflicts with development, construction, or other economic activity given the economic value of older forest which typically is spotted owl habitat. A moderate degree of threat equates to a continual population decline and threat to its habitat, although extinction is not imminent. While the Service is optimistic regarding the potential for recovery, there is uncertainty regarding our ability to alleviate the barred owl impacts to spotted owls and the techniques are still experimental, which matches our guidelines’ “low recovery potential” definition (USDI FWS 1983a p.43101-43104, b p.51985). The spotted owl was originally listed with a recovery priority number of 3C, but that number was changed to 6C in 2004 during the 5-year review of the species (USDI FWS 2004a, p. 55) and to 12C in the 2010 Draft Revised Recovery Plan for the Northern Spotted Owl (USDI FWS 2010, p.22). The revised recovery plan for the northern spotted owl also assigns the spotted owl a recovery priority number of 12C (USDI FWS 2011a p. I-6)

Life History

Taxonomy The northern spotted owl is one of three subspecies of spotted owls currently recognized by the American Ornithologists’ Union. The taxonomic separation of these three subspecies is supported by genetic, (Barrowclough and Gutiérrez 1990, pp.741-742; Barrowclough et al. 1999, p. 928; Haig et al. 2004, p. 1354) morphological (Gutiérrez et al. 1995, p. 2), and biogeographic information (Barrowclough and Gutiérrez 1990, pp.741-742). The distribution of the Mexican subspecies (S. o. lucida) is separate from those of the northern and (S. o. occidentalis) subspecies (Gutiérrez et al. 1995, p.2). Recent studies analyzing mitochondrial DNA sequences (Haig et al. 2004, p. 1354, Chi et al. 2004, p. 3; Barrowclough et al. 2005, p. 1117) and microsatellites (Henke et al., unpubl. data, p. 15) confirmed the validity of the current subspecies designations for northern and California spotted owls. The narrow zone between these two subspecies, which is located in the southern Cascades and northern Sierra Nevada’s, appears to be stable (Barrowclough et al. 2005, p. 1116).

Physical Description The spotted owl is a medium-sized owl and is the largest of the three subspecies of spotted owls (Gutiérrez 1996, p. 2). It is approximately 46 to 48 centimeters (18 inches to 19 inches) long and the sexes are dimorphic, with males averaging about 13 percent smaller than females. The mean mass of 971 males taken during 1,108 captures was 580.4 grams (1.28 pounds) (out of a range 430.0 to 690.0 grams) (0.95 pound to 1.52 pounds), and the mean mass of 874 females taken during 1,016 captures was 664.5 grams (1.46 pounds) (out of a range 490.0 to 885.0 grams) (1.1 pounds to 1.95 pounds) (P. Loschl and E. Forsman, pers. comm. cited in USDI FWS 2008a, p. 43). The spotted owl is dark brown with a barred tail and white spots on its head and breast, and it has dark brown eyes surrounded by prominent facial disks. Four age classes can be distinguished on the basis of plumage characteristics (Moen et al. 1991, p. 493). The spotted owl superficially resembles the barred owl, a species with which 5 it occasionally hybridizes (Kelly and Forsman 2004, p. 807). Hybrids exhibit physical and vocal characteristics of both species (Hamer et al. 1994, p. 488).

Current and Historical Range The current range of the spotted owl extends from southwest through the Cascade Mountains, coastal ranges, and intervening forested lands in , Oregon, and California, as far south as Marin County (USDI FWS 1990, p. 26115). The range of the spotted owl is partitioned into 12 physiographic provinces (Figure 2) based on recognized landscape subdivisions exhibiting different physical and environmental features (USDI FWS 1992a, p. 31). These provinces are distributed across the species’ range as follows (Figure 2): Four provinces in Washington: Eastern Washington Cascades, Olympic Peninsula, Western Washington Cascades, Western Washington Lowlands Five provinces in Oregon: Oregon Coast Range, Willamette Valley, Western Oregon Cascades, Eastern Oregon Cascades, Oregon Klamath Three provinces in California: California Coast, California Klamath, California Cascades

The spotted owl is extirpated or uncommon in certain areas such as southwestern Washington and British Columbia. Timber harvest activities have eliminated, reduced or fragmented spotted owl habitat sufficiently to decrease overall population densities across its range, particularly within the coastal provinces where habitat reduction has been concentrated (USDI FWS 1992b, p. 1799).

Behavior Spotted owls are territorial. However, home ranges of adjacent pairs overlap (Forsman et al. 1984, p. 22; Solis and Gutiérrez 1990, p. 746) suggesting that the area defended is smaller than the area used for foraging. Territorial defense is primarily effected by hooting, barking and whistle type calls. Some spotted owls are not territorial but either remain as residents within the territory of a pair or move among territories (Gutiérrez 1996, p. 4). These are referred to as “floaters.” Floaters have special significance in spotted owl populations because they may buffer the territorial population from decline (Franklin 1992, p. 822). Little is known about floaters other than that they exist and typically do not respond to calls as vigorously as territorial birds (Gutiérrez 1996, p. 4).

Spotted owls are monogamous and usually form long-term pair bonds. “Divorces” occur but are relatively uncommon. There are no known examples of polygyny in this owl, although associations of three or more birds have been reported (Gutiérrez et al. 1995, p. 10).

6

Figure 2. The 12 physiographic Provinces within the range of the northern spotted owl (USDI FWS 1992a, p. 31).

Habitat Relationships Home Range. Home-range sizes vary geographically, generally increasing from south to north, which is likely a response to differences in habitat quality (USDI FWS 1990, p. 26117). Estimates of median size of their annual home range (the area traversed by an individual or pair during their normal activities (Thomas and Raphael 1993, p. IX-15) vary by province and range from 2,955 acres in the Oregon Cascades (Thomas et al. 1990, p. 194) to 14,211 acres on the Olympic Peninsula (USDI FWS 1994a, p. 3). Zabel et al. (1995, p. 436) showed that these provincial home ranges are larger where flying squirrels (Glaucomys sabrinus) are the predominant prey and smaller where wood rats are the predominant prey. Home ranges of adjacent pairs overlap (Forsman et al. 1984, p. 22; Solis and Gutiérrez 1990, p. 746), suggesting that the defended area is smaller than the area used for foraging. Within the home range there is a smaller area of concentrated use during the breeding season (~20% of the home range), often referred to as the core area (Bingham and Noon 1997, pp. 133-135). Spotted owl core areas vary in size geographically and provide habitat elements that are important for the reproductive efficacy of the territory, such as the nest tree, roost sites and foraging areas (Bingham and Noon 1997, p. 134). Spotted owls use smaller home ranges during the breeding season and often dramatically increase their home range size during fall and winter (Forsman et al. 1984, pp. 21-22; Sisco 1990, p. iii).

Although differences exist in natural stand characteristics that influence home range size, habitat loss and forest fragmentation effectively reduce habitat quality in the home range. A reduction in the

7 amount of suitable habitat reduces spotted owl nesting success (Bart 1995, p. 944) and abundance (Bart and Forsman 1992, pp. 98-99).

Habitat Use Forsman et al. (1984, pp.15-16) reported that spotted owls have been observed in the following forest types: Douglas-fir (Pseudotsuga menziesii), western hemlock (Tsuga heterophylla), grand fir (Abies grandis), white fir (Abies concolor), ponderosa pine (Pinus ponderosa), Shasta red fir (Abies magnifica shastensis), mixed evergreen, mixed conifer hardwood (Klamath montane), and redwood (Sequoia sempervirens). The upper elevation limit at which spotted owls occur corresponds to the transition to subalpine forest, which is characterized by relatively simple structure and severe winter weather (Forsman 1975, p. 27; Forsman et al. 1984, pp. 15-16).

Roost sites selected by spotted owls have more complex vegetation structure than forests generally available to them (Barrows and Barrows 1978, p.3; Forsman et al. 1984, pp.29-30; Solis and Gutiérrez 1990, pp.742-743). These habitats are usually multi-layered forests having high canopy closure and large diameter trees in the overstory.

Spotted owls nest almost exclusively in trees. Like roosts, nest sites are found in forests having complex structure dominated by large diameter trees (Forsman et al. 1984, p.30; Hershey et al. 1998, p.1402). Even in forests that have been previously logged, spotted owls select forests having a structure (i.e., larger trees, greater canopy closure) different than forests generally available to them (Folliard 1993, p. 40; Buchanan et al. 1995, p.1402; Hershey et al. 1998 p. 1404).

Foraging habitat is the most variable of all habitats used by territorial spotted owls (USDI FWS 1992a, p. 20). Descriptions of foraging habitat have ranged from complex structure (Solis and Gutiérrez 1990, pp. 742-744) to forests with lower canopy closure and smaller trees than forests containing nests or roosts (Gutiérrez 1996, p.5).

Habitat Selection Spotted owls generally rely on older forested habitats because such forests contain the structures and characteristics required for nesting, roosting, and foraging. Features that support nesting and roosting typically include a moderate to high canopy closure (60 to 90 percent); a multi-layered, multi-species canopy with large overstory trees (with dbh of greater than 30 inches); a high incidence of large trees with various deformities (large cavities, broken tops, mistletoe infections, and other evidence of decadence); large snags; large accumulations of fallen trees and other woody debris on the ground; and sufficient open space below the canopy for spotted owls to fly (Thomas et al. 1990, p. 19). Nesting spotted owls consistently occupy stands with a high degree of canopy closure that may provide thermoregulatory benefits (Weathers et al. 2001, p. 686) and protection from predators.

Foraging habitat for spotted owls provides a food supply for survival and reproduction. Foraging activity is positively associated with tree height diversity (North et al. 1999, p. 524), canopy closure (Irwin et al. 2000, p. 180; Courtney et al. 2004, p. 5-15), snag volume, density of snags greater than 20 in (50 cm) dbh (North et al. 1999, p. 524; Irwin et al. 2000, pp. 179-180; Courtney et al. 2004, p. 5-15), density of trees greater than or equal to 31 in (80 cm) dbh (North et al. 1999, p. 524), volume of woody debris (Irwin et al. 2000, pp. 179-180), and young forests with some structural characteristics of old forests (Carey et al.1992, pp. 245-247; Irwin et al. 2000, pp. 178-179). Spotted owls select old forests for foraging in greater proportion than their availability at the landscape scale (Carey et al. 1992, pp. 236- 237; Carey and Peeler 1995, p. 235; Forsman et al. 2005, pp. 372-373), but will forage in younger stands

8 with high prey densities and access to prey (Carey et al. 1992, p. 247; Rosenberg and Anthony 1992, p. 165; Thome et al. 1999, p. 56-57).

Dispersal habitat is essential to maintaining stable populations by filling territorial vacancies when resident spotted owls die or leave their territories, and to providing adequate gene flow across the range of the species. Dispersal habitat, at a minimum, consists of stands with adequate tree size and canopy closure to provide protection from avian predators and at least minimal foraging opportunities. Dispersal habitat may include younger and less diverse forest stands than foraging habitat, such as even- aged, pole-sized stands, but such stands should contain some roosting structures and foraging habitat to allow for temporary resting and feeding for dispersing juveniles (USDI FWS 1992b, p. 1798). Forsman et al. (2002, p. 22) found that spotted owls could disperse through highly fragmented forest landscapes. However, the stand-level and landscape-level attributes of forests needed to facilitate successful dispersal have not been thoroughly evaluated (Buchanan 2004, p. 1341).

Spotted owls may be found in younger forest stands that have the structural characteristics of older forests or retained structural elements from the previous forest. In redwood forests and mixed conifer- hardwood forests along the coast of northwestern California, considerable numbers of spotted owls also occur in younger forest stands, particularly in areas where hardwoods provide a multi-layered structure at an early age (Thomas et al. 1990, p. 158; Diller and Thome 1999, p. 275). In mixed conifer forests in the eastern Cascades in Washington, 27 percent of nest sites were in old-growth forests, 57 percent were in the understory reinitiation phase of stand development, and 17 percent were in the stem exclusion phase (Buchanan et al. 1995, p. 304). In the western Cascades of Oregon, 50 percent of spotted owl nests were in late-seral/old-growth stands (greater than 80 years old), and none were found in stands of less than 40 years old (Irwin et al. 2000, p. 41).

In the Western Washington Cascades, spotted owls roosted in mature forests dominated by trees greater than 50 centimeters (19.7 inches) dbh with greater than 60 percent canopy closure more often than expected for roosting during the non-breeding season. Spotted owls also used young forest (trees of 20 to 50 centimeters (7.9 inches to 19.7 inches) dbh with greater than 60 percent canopy closure) less often than expected based on this habitat’s availability (Herter et al. 2002, p. 437).

In the Coast Ranges, Western Oregon Cascades and the Olympic Peninsula, radio-marked spotted owls selected for old-growth and mature forests for foraging and roosting and used young forests less than predicted based on availability (Forsman et al. 1984, pp. 24-25; Carey et al. 1990, pp. 14-15; Forsman et al. 2005, pp. 372-373). Glenn et al. (2004, pp. 46-47) studied spotted owls in young forests in western Oregon and found little preference among age classes of young forest with spotted owls generally selecting mature and older conifer and broadleaf forest types.

Habitat use is influenced by prey availability. Ward (1990, p. 62) found that spotted owls foraged in areas with lower variance in prey densities (that is, where the occurrence of prey was more predictable) within older forests and near ecotones of old forest and brush seral stages. Zabel et al. (1995, p. 436) showed that spotted owl home ranges are larger where flying squirrels (Glaucomys sabrinus) are the predominant prey and smaller where wood rats (Neotoma spp.) are the predominant prey.

Recent landscape-level analyses in portions of Oregon Coast and California Klamath provinces suggest that a mosaic of late-successional habitat interspersed with other seral conditions may benefit spotted owls more than large, homogeneous expanses of older forests (Zabel et al. 2003, p. 1038; Franklin et al. 2000, pp. 573-579; Meyer et al. 1998, p. 43). In Oregon Klamath and Western Oregon Cascade

9 provinces, Dugger et al. (2005, p. 876) found that apparent survival and reproduction was positively associated with the proportion of older forest near the territory center (within 730 meters) (2,395 feet). Survival decreased dramatically when the amount of non-habitat (non-forest areas, sapling stands, etc.) exceeded approximately 50 percent of the home range (Dugger et al. 2005, pp. 873-874). The authors concluded that they found no support for either a positive or negative direct effect of intermediate-aged forest—that is, all forest stages between sapling and mature, with total canopy cover greater than 40 percent—on either the survival or reproduction of spotted owls. It is unknown how these results were affected by the low habitat fitness potential in their study area, which Dugger et al. (2005, p. 876) stated was generally much lower than those in Franklin et al. (2000) and Olson et al. (2004), and the low reproductive rate and survival in their study area, which they reported were generally lower than those studied by Anthony et al. (2006). Olson et al. (2004, pp. 1050-1051) found that reproductive rates fluctuated biennially and were positively related to the amount of edge between late-seral and mid- seral forests and other habitat classes in the central Oregon Coast Range. Olson et al. (2004, pp. 1049- 1050) concluded that their results indicate that while mid-seral and late-seral forests are important to spotted owls, a mixture of these forest types with younger forest and non-forest may be best for spotted owl survival and reproduction in their study area.

Reproductive Biology The spotted owl is relatively long-lived, has a long reproductive life span, invests significantly in parental care, and exhibits high adult survivorship relative to other North American owls (Gutiérrez et al. 1995, p. 5). Spotted owls are sexually mature at 1 year of age, but rarely breed until they are 2 to 5 years of age (Miller et al. 1985, p. 93; Franklin 1992, p. 821; Forsman et al. 2002, p. 17). Breeding females lay one to four eggs per clutch, with the average clutch size being two eggs; however, most spotted owl pairs do not nest every year, nor are nesting pairs successful every year (Forsman et al. 1984, pp. 32-34, Anthony et al. 2006, p. 28), and renesting after a failed nesting attempt is rare (Gutiérrez 1996, p. 4). The small clutch size, temporal variability in nesting success, and delayed onset of breeding all contribute to the relatively low fecundity of this species (Gutiérrez 1996, p. 4).

Courtship behavior usually begins in February or March, and females typically lay eggs in late March or April. The timing of nesting and fledging varies with latitude and elevation (Forsman et al. 1984, p. 32). After they leave the nest in late May or June, juvenile spotted owls depend on their parents until they are able to fly and hunt on their own. Parental care continues after fledging into September (Forsman et al. 1984, p. 38). During the first few weeks after the young leave the nest, the adults often roost with them during the day. By late summer, the adults are rarely found roosting with their young and usually only visit the juveniles to feed them at night (Forsman et al. 1984, p. 38). Telemetry and genetic studies indicate that close inbreeding between siblings or parents and their offspring is rare (Haig et al. 2001, p. 35, Forsman et al. 2002, p. 18).

Dispersal Biology Natal dispersal of spotted owls typically occurs in September and October with a few individuals dispersing in November and December (Forsman et al. 2002, p. 13). Natal dispersal occurs in stages, with juveniles settling in temporary home ranges between bouts of dispersal (Forsman et al. 2002, pp. 13-14; Miller et al. 1997, p. 143). The median natal dispersal distance is about 10 miles for males and 15.5 miles for females (Forsman et al. 2002, p. 16). Dispersing juvenile spotted owls experience high mortality rates, exceeding 70 percent in some studies (Miller 1989, pp. 32-41). Known or suspected causes of mortality during dispersal include starvation, predation, and accidents (Miller 1989, pp. 41-44; Forsman et al. 2002, pp. 18-19). Parasitic infection may contribute to these causes of mortality, but the relationship between parasite loads and survival is poorly understood (Hoberg et al. 1989, p. 247;

10

Gutiérrez 1989, pp. 616-617, Forsman et al. 2002, pp. 18-19). Successful dispersal of juvenile spotted owls may depend on their ability to locate unoccupied suitable habitat in close proximity to other occupied sites (LaHaye et al. 2001, pp. 697-698).

There is little evidence that small openings in forest habitat influence the dispersal of spotted owls, but large, non-forested valleys such as the Willamette Valley apparently are barriers to both natal and breeding dispersal (Forsman et al. 2002, p. 22). The degree to which water bodies, such as the Columbia River and Puget Sound, function as barriers to dispersal is unclear, although radio telemetry data indicate that spotted owls move around large water bodies rather than cross them (Forsman et al. 2002, p. 22). Analysis of the genetic structure of spotted owl populations suggests that gene flow may have been adequate between the Olympic Mountains and the Washington Cascades, and between the Olympic Mountains and the Oregon Coast Range (Haig et al. 2001, p. 35).

Breeding dispersal occurs among a small proportion of adult spotted owls; these movements were more frequent among females and unmated individuals (Forsman et al. 2002, pp. 20-21). Breeding dispersal distances were shorter than natal dispersal distances and also are apparently random in direction (Forsman et al. 2002, pp. 21-22).

Food Habits Spotted owls are mostly nocturnal, although they also forage opportunistically during the day (Forsman et al. 1984, p. 51; 2004, pp. 222-223; Sovern et al. 1994, p. 202). The composition of the spotted owl’s diet varies geographically and by forest type. Generally, flying squirrels (Glaucomys sabrinus) are the most prominent prey for spotted owls in Douglas-fir and western hemlock (Tsuga heterophylla) forests (Forsman et al. 1984, pp. 40-41) in Washington (Hamer et al. 2001, p. 224) and Oregon, while dusky- footed wood rats (Neotoma fuscipes) are a major part of the diet in the Oregon Klamath, California Klamath, and California Coastal provinces (Forsman et al. 1984, pp. 40-42; 2004, p. 218; Ward et al. 1998, p. 84). Depending on location, other important prey include deer mice (Peromyscus maniculatus), tree (Arborimus longicaudus, A. pomo), red-backed voles (Clethrionomys spp.), (Thomomys spp.), snowshoe hare (Lepus americanus), bushy-tailed wood rats (Neotoma cinerea), birds, and insects, although these species comprise a small portion of the spotted owl diet (Forsman et al. 1984, pp. 40-43; 2004, p. 218; Ward et al. 1998; p. 84; Hamer et al. 2001, p.224).

Other prey species such as the red tree (Arborimus longicaudus), red-backed voles (Clethrionomys gapperi), mice, rabbits and hares, birds, and insects) may be seasonally or locally important (reviewed by Courtney et al. 2004, p. 4-27). For example, Rosenberg et al. (2003, p. 1720) showed a strong correlation between annual reproductive success of spotted owls (number of young per territory) and abundance of deer mice (Peromyscus maniculatus) (r2 = 0.68), despite the fact they only made up 1.6±0.5 percent of the biomass consumed. However, it is unclear if the causative factor behind this correlation was prey abundance or a synergistic response to weather (Rosenberg et al. 2003, p. 1723). Ward (1990, p. 55) also noted that mice were more abundant in areas selected for foraging by owls. Nonetheless, spotted owls deliver larger prey to the nest and eat smaller food items to reduce foraging energy costs; therefore, the importance of smaller prey items, like Peromyscus, in the spotted owl diet should not be underestimated (Forsman et al. 2001, p. 148; 2004, pp. 218-219).

Population Dynamics The spotted owl is relatively long-lived, has a long reproductive life span, invests significantly in parental care, and exhibits high adult survivorship relative to other North American owls (Gutiérrez 1996, p. 5).

11

The spotted owl’s long reproductive life span allows for some eventual recruitment of offspring, even if recruitment does not occur each year (Franklin et al. 2000, p. 576).

Annual variation in population parameters for spotted owls has been linked to environmental influences at various life history stages (Franklin et al. 2000, p. 581). In coniferous forests, mean fledgling production of the California spotted owl (Strix occidentalis occidentalis), a closely related subspecies, was higher when minimum spring temperatures were higher (North et al. 2000, p. 805), a relationship that may be a function of increased prey availability. Across their range, spotted owls have previously shown an unexplained pattern of alternating years of high and low reproduction, with highest reproduction occurring during even-numbered years (e.g., Franklin et al. 1999, p. 1). Annual variation in breeding may be related to weather (i.e., temperature and precipitation) (Wagner et al. 1996, p. 74 and Zabel et al. 1996, p.81 In: Forsman et al. 1996) and fluctuation in prey abundance (Zabel et al. 1996, p.437-438).

A variety of factors may regulate spotted owl population levels. These factors may be density- dependent (e.g., habitat quality, habitat abundance) or density-independent (e.g., climate). Interactions may occur among factors. For example, as habitat quality decreases, density-independent factors may have more influence on survival and reproduction, which tends to increase variation in the rate of growth (Franklin et al. 2000, pp. 581-582). Specifically, weather could have increased negative effects on spotted owl fitness for those owls occurring in relatively lower quality habitat (Franklin et al. 2000, pp. 581-582). A consequence of this pattern is that at some point, lower habitat quality may cause the population to be unregulated (have negative growth) and decline to extinction (Franklin et al. 2000, p. 583).

Olson et al. (2005, pp. 930-931) used open population modeling of site occupancy that incorporated imperfect and variable detectability of spotted owls and allowed modeling of temporal variation in site occupancy, extinction, and colonization probabilities (at the site scale). The authors found that visit detection probabilities average less than 0.70 and were highly variable among study years and among their three study areas in Oregon. Pair site occupancy probabilities declined greatly on one study area and slightly on the other two areas. However, for all owls, including singles and pairs, site occupancy was mostly stable through time. Barred owl presence had a negative effect on these parameters (see barred owl discussion in the New Threats section below). However, there was enough temporal and spatial variability in detection rates to indicate that more visits would be needed in some years and in some areas, especially if establishing pair occupancy was the primary goal.

Threats

Reasons for Listing The spotted owl was listed as threatened throughout its range “due to loss and adverse modification of suitable habitat as a result of timber harvesting and exacerbated by catastrophic events such as fire, volcanic eruption, and wind storms” (USDI FWS 1990, p. 26114). More specifically, threats to the spotted owl included low populations, declining populations, limited habitat, declining habitat, inadequate distribution of habitat or populations, isolation of provinces, predation and competition, lack of coordinated conservation measures, and vulnerability to natural disturbance (USDI FWS 1992b, pp. 33-41). These threats were characterized for each province as severe, moderate, low or unknown (USDI FWS 1992b, p. 33-41) (The range of the spotted owl is divided into 12 provinces from to northern California and from the Pacific Coast to the eastern Cascades; see Figure 3). Declining habitat was recognized as a severe or moderate threat to the spotted owl throughout its range, isolation of

12 populations was identified as a severe or moderate threat in 11 provinces, and a decline in population was a severe or moderate threat in 10 provinces. Together, these three factors represented the greatest concerns about range-wide conservation of the spotted owl. Limited habitat was considered a severe or moderate threat in nine provinces, and low populations were a severe or moderate concern in eight provinces, suggesting that these factors were also a concern throughout the majority of the spotted owl’s range. Vulnerability to natural disturbances was rated as low in five provinces.

The degree to which predation and competition might pose a threat to the spotted owl was unknown in more provinces than any of the other threats, indicating a need for additional information. Few empirical studies exist to confirm that habitat fragmentation contributes to increased levels of predation on spotted owls (Courtney et al. 2004, pp. 11-8 to 11-9). However, great horned owls (Bubo virginianus), an effective predator on spotted owls, are closely associated with fragmented forests, openings, and clearcuts (Johnson 1992, p. 84; Laidig and Dobkin 1995, p. 155). As mature forests are harvested, great horned owls may colonize fragmented forests, thereby increasing spotted owl vulnerability to predation.

New Threats The Service conducted a 5-year review of the spotted owl in 1994 (USDI FWS 2004a), for which the Service prepared a scientific evaluation of the status of the spotted owl (Courtney et al. 2004). An analysis was conducted assessing how the threats described in 1990 might have changed by 2004. Some of the key threats identified in 2004 are: “Although we are certain that current harvest effects are reduced, and that past harvest is also probably having a reduced effect now as compared to 1990, we are still unable to fully evaluate the current levels of threat posed by harvest because of the potential for lag effects…In their questionnaire responses…6 of 8 panel member identified past habitat loss due to timber harvest as a current threat, but only 4 viewed current harvest as a present threat” (Courtney and Gutiérrez 2004, p. 11-7) “Currently the primary source of habitat loss is catastrophic wildfire, although the total amount of habitat affected by wildfires has been small (a total of 2.3% of the range-wide habitat base over a 10- year period).” (Courtney and Gutiérrez 2004, p. 11-8) “Although the panel had strong differences of opinion on the conclusiveness of some of the evidence suggesting [barred owl] displacement of [spotted owls], and the mechanisms by which this might be occurring, there was no disagreement that [barred owls] represented an operational threat. In the questionnaire, all 8 panel members identified [barred owls] as a current threat, and also expressed concern about future trends in [barred owl] populations.” (Courtney and Gutiérrez 2004, p. 11-8)

Barred Owls (Strix varia) With its recent expansion to as far south as Marin County, California along the Coast Range and Kings Canyon National Park in the southern Sierra Nevada (Gutiérrez et al. 2004, pp. 7-12-7-13; Steger et al.2006), the barred owl’s range now completely overlaps that of the spotted owl. Barred owls may be competing with spotted owls for prey (Hamer et al. 2001, p.226) or habitat (Hamer et al. 1989, p.55; Dunbar et al. 1991, p. 467; Herter and Hicks 2000, p. 285; Pearson and Livezey 2003, p. 274; Wiens 2012). In addition, barred owls physically attack spotted owls (Pearson and Livezey 2003, p. 274), and circumstantial evidence strongly indicated that a barred owl killed a spotted owl (Leskiw and Gutiérrez 1998, p. 226). Evidence that barred owls are causing negative effects on spotted owls is largely indirect, based primarily on retrospective examination of long-term data collected on spotted owls (Kelly et al. 2003, p. 46; Pearson and Livezey 2003, p. 267; Olson et al. 2005, p. 921). The two species of owls appear to be competing for resources (Hamer et al.2001 p 226, Wiens 2012). However, given that the presence of barred owls has been identified as a negative effect while using methods designed to detect a

13 different species (spotted owls), it is likely the effects are stronger than estimated (Wiens et al.2011). Because there has been no research to quantitatively evaluate the strength of different types of competitive interactions, such as resource partitioning and competitive interference, the particular mechanism by which the two owl species may be competing is unknown.

Barred owls were initially thought to be more closely associated with early successional forests than spotted owls, based on studies conducted on the west slope of the Cascades in Washington (Hamer et al.1989, p. 34; Iverson 1993, p.39). However, recent studies conducted in the show that barred owls frequently use mature and old-growth forests (Pearson and Livezey 2003, p. 270; Schmidt 2006, p. 13, Wiens 2012 p. 65). In the fire prone forests of eastern Washington, a telemetry study conducted on barred owls showed that barred owl home ranges were located on lower slopes or valley bottoms, in closed canopy, mature, Douglas-fir forest, while spotted owl sites were located on mid-elevation areas with southern or western exposure, characterized by closed canopy, mature, ponderosa pine or Douglas-fir forest (Singleton et al. 2005, p. 1).

Barred owls, though they are generalists, likely compete with spotted owls for prey resources (Hamer et al.2001, p 225; Gutierrez et al.p. 187; Livezy and Fleming 2007, p. 319). In one study comparing spotted owl and barred owl food habits in the Pacific Northwest, Hamer et al. (2001, p. 221, 226) indicated barred owl diets overlap strongly (76 percent) with spotted owl diets. Wiens et al.(2012 p. 3) found dietary overlap between pairs of spotted and barred owls in adjacent territories ranged from 28–70% (mean = 42%). However, barred owl diets are more diverse than spotted owl diets and include species associated with riparian and other moist habitats (e.g. fish, invertebrates, frogs, and crayfish), along with more terrestrial and diurnal species (Hamer et al. 2001, pp. 225-226; Wiens et al. 2012 p. 3)). Even though barred owls may be taking spotted owls’ primary prey only as a generalist, spotted owls may be affected by a reduction in the density of these prey. This could lead to depletion of available spotted owl prey to the extent that spotted owls cannot survive or reproduce (Gutierrez et al.2007 p. 187; Livezy and Fleming 2007, p. 319; Wiens 2012 p. 57).

The presence of barred owls has been reported to reduce spotted owl detectability, site occupancy, reproduction, and survival. Olson et al. (2005, p. 924) found that the presence of barred owls had a significant negative effect on the detectability of spotted owls, and that the magnitude of this effect did not vary among years. The occupancy of historical territories by spotted owls in Washington and Oregon was significantly lower (p < 0.001) after barred owls were detected within 0.8 kilometer (0.5 miles) of the territory center but was “only marginally lower” (p = 0.06) if barred owls were located more than 0.8 kilometer (0.5 miles) from the spotted owl territory center (Kelly et al. 2003, p. 51). Pearson and Livezey (2003, p. 271) found significantly more barred owl site-centers in unoccupied spotted owl circles than occupied spotted owl circles (centered on historical spotted owl site-centers) with radii of 0.8 kilometer (0.5 miles) (p = 0.001), 1.6 kilometer (1 mile) (p = 0.049), and 2.9 kilometer (1.8 miles) (p = 0.005) in Gifford Pinchot National Forest. In , Gremel (2005, p. 11) found a significant decline (p = 0.01) in spotted owl pair occupancy at sites where barred owls had been detected, while pair occupancy remained stable at spotted owl sites without barred owls. Olson et al. (2005, p. 928) found that the annual probability that a spotted owl territory would be occupied by a pair of spotted owls after barred owls were detected at the site declined by 5 percent in the HJ Andrews study area, 12 percent in the Coast Range study area, and 15 percent in the Tyee study area.

In contrast, Bailey et al. (2009, p. 2983), when using a two-species occupancy model, showed no evidence that barred owls exclude spotted owls from territories in Oregon. Most recently, preliminary results from a barred owl and spotted owl radio-telemetry study in Washington reported two spotted

14 owls fleeing their territories and traveling six and 15 miles, believed to be as a result of frequent direct encounters with barred owls (Irwin et al. 2010 pp. 3-4). Both spotted owls were subsequently found dead (Irwin et al. 2010, p. 4).

Olson et al. (2004, p. 1048) found that the presence of barred owls had a significant negative effect on the reproduction of spotted owls in the central Coast Range of Oregon (in the Roseburg study area). The conclusion that barred owls had no significant effect on the reproduction of spotted owls in one study (Iverson 2004, p. 89) was unfounded because of small sample sizes (Livezey 2005, p. 102). It is likely that all of the above analyses underestimated the effects of barred owls on the reproduction of spotted owls because spotted owls often cannot be relocated after they are displaced by barred owls (E. Forsman, pers. comm., cited in USDI FWS 2011a, p. B-11). Forsman et al. (2011) documented negative effects of barred owls on apparent survival at 5 study areas and negative effects of barred owls on fecundity at 4 of 11 study areas. They attributed the equivocal results for most of their study areas to the coarse nature of their barred owl covariate. Dugger et al.(2011, pp. 2463-2467) confirmed the synergistic effects of barred owls and territory habitat characteristics on extinction and colonization rates of territories by spotted owls. Extinction rates of spotted owl territories nearly tripled when barred owls were detected (Dugger et al.2011 p. 2464).

Monitoring and management of northern spotted owls has become more complicated due to their possible reduced detectability when barred owls are present (Kelly et al. 2003, pp. 51-52; Courtney et al. 2004, p. 7-16 ; Olson et al. 2005, p. 929; Crozier et al. 2006, p.766-767). Evidence that northern spotted owls were responding less frequently during surveys led the Service and its many research partners to update the northern spotted owl survey protocol. The recent changes to the northern spotted owl survey protocol were based on the probability of detecting northern spotted owls when barred owls are present (USDI FWS 2011b).

In a recent analysis of more than 9,000 banded spotted owls throughout their range, only 47 hybrids were detected (Kelly and Forsman 2004, p. 807). Consequently, hybridization with the barred owl is considered to be “an interesting biological phenomenon that is probably inconsequential, compared with the real threat—direct competition between the two species for food and space” (Kelly and Forsman 2004, p. 808).

The preponderance of evidence suggests that barred owls are exacerbating the spotted owl population decline, particularly in Washington, portions of Oregon, and the northern coast of California (Gutiérrez et al. 2004, pp. 739-740; Olson et al. 2005, pp. 930-931). There is no evidence that the increasing trend in barred owls has stabilized in any portion of the spotted owl’s range in the western United States, and “there are no grounds for optimistic views suggesting that barred owl impacts on spotted owls have been already fully realized” (Gutiérrez et al. 2004, pp. 7-38).

Wildfire Studies indicate that the effects of wildfire on spotted owls and their habitat are variable, depending on fire intensity, severity and size. Within the fire-adapted forests of the spotted owl’s range, spotted owls likely have adapted to withstand fires of variable sizes and severities. Bond et al. (2002, p. 1025) examined the demography of the three spotted owl subspecies after wildfires, in which wildfire burned through spotted owl nest and roost sites in varying degrees of severity. Post-fire demography parameters for the three subspecies were similar or better than long-term demographic parameters for each of the three subspecies in those same areas (Bond et al. 2002, p. 1026). In a preliminary study conducted by Anthony and Andrews (2004, p. 8) in the Oregon Klamath Province, their sample of

15 spotted owls appeared to be using a variety of habitats within the area of the Timbered Rock fire, including areas where burning had been moderate.

In 1994, the Hatchery Complex fire burned 17,603 hectares in the Wenatchee National Forest in Washington’s eastern Cascades, affecting six spotted owl activity centers (Gaines et al. 1997, p. 125). Spotted owl habitat within a 2.9-kilometer (1.8-mile) radius of the activity centers was reduced by 8 to 45 percent (mean = 31 percent) as a result of the direct effects of the fire and by 10 to 85 percent (mean = 55 percent) as a result of delayed mortality of fire-damaged trees and insects. Direct mortality of spotted owls was assumed to have occurred at one site, and spotted owls were present at only one of the six sites 1 year after the fire (Gaines et al. 1997, p. 126). In 1994, two wildfires burned in the Yakama Indian Reservation in Washington’s eastern Cascades, affecting the home ranges of two radio-tagged spotted owls (King et al. 1998, pp. 2-3). Although the amount of home ranges burned was not quantified, spotted owls were observed using areas that burned at low and medium intensities. No direct mortality of spotted owls was observed, even though thick smoke covered several spotted owl site-centers for a week. It appears that, at least in the short term, spotted owls may be resilient to the effects of wildfire—a process with which they have evolved. More research is needed to further understand the relationship between fire and spotted owl habitat use.

At the time of listing there was recognition that large-scale wildfire posed a threat to the spotted owl and its habitat (USDI FWS 1990, p. 26183). New information suggests fire may be more of a threat than previously thought. In particular, the rate of habitat loss due to fire has been expected with over 102,000 acres of late-successional forest lost on Federal lands from 1993-2004 (Moeur et al. 2005, p. 110). Currently, the overall total amount of habitat loss from wildfires has been relatively small, estimated at approximately 1.2 percent on federal lands (Lint et al. 2005, p. v). It may be possible to influence through silvicultural management how fire prone forests will burn and the extent of the fire when it occurs. Silvicultural management of forest fuels are currently being implemented throughout the spotted owl’s range, in an attempt to reduce the levels of fuels that have accumulated during nearly 100 years of effective fire suppression. However, our ability to protect spotted owl habitat and viable populations of spotted owls from large fires through risk-reduction endeavors is uncertain (Courtney et al. 2004, pp. 12-11). The Northwest Forest Plan (NWFP) recognized wildfire as an inherent part of managing spotted owl habitat in certain portions of the range. The distribution and size of reserve blocks as part of the NWFP design may help mitigate the risks associated with large-scale fire (Lint et al. 2005, p. 77).

West Nile Virus (WNV) WNV has killed millions of wild birds in since it arrived in 1999 (Caffrey and Peterson 2003, p. 12; Marra et al. 2004, p. 393). Mosquitoes are the primary carriers (vectors) of the virus that causes encephalitis in humans, horses, and birds. Mammalian prey may also play a role in spreading WNV among predators, like spotted owls. Owls and other predators of mice can contract the disease by eating infected prey (Garmendia et al. 2000, p. 3111). One captive spotted owl in Ontario, Canada, is known to have contracted WNV and died (Gancz et al. 2004, p. 2137), but there are no documented cases of the virus in wild spotted owls.

Health officials expect that WNV eventually will spread throughout the range of the spotted owl (Blakesley et al. 2004, p. 8-31), but it is unknown how the virus will ultimately affect spotted owl populations. Susceptibility to infection and the mortality rates of infected individuals vary among species (Blakesley et al. 2004, p. 8-33), but most owls appear to be quite susceptible. For example, eastern screech-owls breeding in Ohio that were exposed to WNV experienced 100 percent mortality (T.

16

Grubb pers. comm. in Blakesley et al. 2004, p. 8-33). Barred owls, in contrast, showed lower susceptibility (B. Hunter pers. comm. in Blakesley et al. 2004, p. 8-34).

Blakesley et al. (2004, p. 8-35) offer two possible scenarios for the likely outcome of spotted owl populations being infected by WNV. One scenario is that a range-wide reduction in spotted owl population viability is unlikely because the risk of contracting WNV varies between regions. An alternative scenario is that WNV will cause unsustainable mortality, due to the frequency and/or magnitude of infection, thereby resulting in long-term population declines and extirpation from parts of the spotted owl’s current range. WNV remains a potential threat of uncertain magnitude and effect (Blakesley et al. 2004, p. 8-34).

Sudden Oak Death Sudden oak death was recently identified as a potential threat to the spotted owl (Courtney and Guttierez 2004, p. 11-8). This disease is caused by the fungus-like pathogen, Phytopthora ramorum that was recently introduced from Europe and is rapidly spreading. At the present time, sudden oak death is found in natural stands from Monterey to Humboldt Counties, California, and has reached epidemic proportions in oak (Quercus spp.) and tanoak (Lithocarpus densiflorus) forests along approximately 300 km of the central and northern California coast (Rizzo et al. 2002, p. 733). It has also been found near Brookings, Oregon, killing tanoak and causing dieback of closely associated wild rhododendron (Rhododendron spp.) and evergreen huckleberry (Vaccinium ovatum) (Goheen et al. 2002, p. 441). It has been found in several different forest types and at elevations from sea level to over 800 m. Sudden oak death poses a threat of uncertain proportion because of its potential impact on forest dynamics and alteration of key prey and spotted owl habitat components (e.g., hardwood trees - canopy closure and nest tree mortality); especially in the southern portion of the spotted owl’s range (Courtney and Guttierez 2004, p. 11-8).

Inbreeding Depression, Genetic Isolation, and Reduced Genetic Diversity Inbreeding and other genetic problems due to small population sizes were not considered an imminent threat to the spotted owl at the time of listing. Recent studies show no indication of significantly reduced genetic variation in Washington, Oregon, or California (Barrowclough et al. 1999, p. 922; Haig et al. 2001, p. 36). However, in Canada, the breeding population is estimated to be less than 33 pairs and annual population decline may be as high as 35 percent (Harestad et al. 2004, p. 13). Canadian populations may be more adversely affected by issues related to small population size including inbreeding depression, genetic isolation, and reduced genetic diversity (Courtney et al. 2004, p. 11-9). Low and persistently declining populations throughout the northern portion of the species range (see “Population Trends” below) may be at increased risk of losing genetic diversity. Climate change. Climate change, a potential additional threat to spotted owl populations, is not explicitly addressed in the NWFP. Climate change could have direct and indirect impacts on spotted owls and their prey. However, the emphasis on maintenance of seral stage complexity and related organismal diversity in the Matrix under the NWFP should contribute to the resiliency of the Federal forest landscape to the impacts of climate change (Courtney et al. 2004, p. 9-15). There is no indication in the literature regarding the direction (positive or negative) of the threat.

Based upon a global meta-analysis, Parmesan and Yohe (2003, pp. 37-42) discussed several potential implications of global climate change to biological systems, including terrestrial flora and fauna. Results indicated that 62 percent of species exhibited trends indicative of advancement of spring conditions. In bird species, trends were manifested in earlier nesting activities. Because the spotted owl exhibits a

17 limited tolerance to heat relative to other bird species (Weathers et al. 2001, p. 685), subtle changes in climate have the potential to affect this. However, the specific impacts to the species are unknown.

Disturbance-Related Effects The effects of noise on spotted owls are largely unknown, and whether noise is a concern has been a controversial issue. The effect of noise on birds is extremely difficult to determine due to the inability of most studies to quantify one or more of the following variables: 1) timing of the disturbance in relation to nesting chronology; 2) type, frequency, and proximity of human disturbance; 3) clutch size; 4) health of individual birds; 5) food supply; and 6) outcome of previous interactions between birds and humans (Knight and Skagan 1988, pp. 355-358). Additional factors that confound the issue of disturbance include the individual bird’s tolerance level, ambient sound levels, physical parameters of sound and how it reacts with topographic characteristics and vegetation, and differences in how species perceive noise.

Although information specific to behavioral responses of spotted owls to disturbance is limited, research indicates that close proximity to recreational hikers can cause Mexican spotted owls (S. o. lucida) to flush from their roosts (Swarthout and Steidl 2001, p. 314) and helicopter overflights can reduce prey delivery rates to nests (Delaney et al. 1999, p. 70). Additional effects from disturbance, including altered foraging behavior and decreases in nest attendance and reproductive success, have been reported for other raptors (White and Thurow 1985, p. 14; Andersen et al. 1989, p. 296; McGarigal et al. 1991, p. 5).

Spotted owls may also respond physiologically to a disturbance without exhibiting a significant behavioral response. In response to environmental stressors, vertebrates secrete stress hormones called corticosteroids (Campbell 1990, p. 925). Although these hormones are essential for survival, extended periods with elevated stress hormone levels may have negative effects on reproductive function, disease resistance, or physical condition (Carsia and Harvey 2000, pp. 517-518; Sapolsky et al. 2000, p. 1). In avian species, the secretion of corticosterone is the primary non-specific stress response (Carsia and Harvey 2000, p. 517). The quantity of this hormone in feces can be used as a measure of physiological stress (Wasser et al.1997, p. 1019). Recent studies of fecal corticosterone levels of spotted owls indicate that low intensity noise of short duration and minimal repetition does not elicit a physiological stress response (Tempel & Gutiérrez 2003, p. 698; Tempel & Gutiérrez 2004, p. 538). However, prolonged activities, such as those associated with timber harvest, may increase fecal corticosterone levels depending on their proximity to spotted owl core areas (Wasser et al. 1997, p.1021; Tempel & Gutiérrez 2004, p. 544).

Post-harvest fuels treatments may also create above-ambient smoke or heat. Although it has not been conclusively demonstrated, it is anticipated that nesting spotted owls may be disturbed by heat and smoke intrusion into the nest grove.

Conservation Needs of the Spotted Owl

Based on the above assessment of threats, the spotted owl has the following habitat-specific and habitat-independent conservation (i.e., survival and recovery) needs:

Habitat-specific Needs 1. Large blocks of suitable habitat to support clusters or local population centers of spotted owls (e.g., 15 to 20 breeding pairs) throughout the owl’s range;

18

2. Suitable habitat conditions and spacing between local spotted owl populations throughout its range to facilitate survival and movement;

3. Suitable habitat distributed across a variety of ecological conditions within the spotted owl’s range to reduce risk of local or widespread extirpation;

4. A coordinated, adaptive management effort to reduce the loss of habitat due to catastrophic wildfire throughout the spotted owl’s range, and a monitoring program to clarify whether these risk reduction methods are effective and to determine how owls use habitat treated to reduce fuels; and

5. In areas of significant population decline, sustain the full range of survival and recovery options for this species in light of significant uncertainty.

Habitat-independent Needs 1. A coordinated research and adaptive management effort to better understand and manage competitive interactions between spotted and barred owls; and

2. Monitoring to better understand the risk that WNV and sudden oak death pose to spotted owls and, for WNV, research into methods that may reduce the likelihood or severity of outbreaks in spotted owl populations.

Conservation Strategy Since 1990, various efforts have addressed the conservation needs of the spotted owl and attempted to formulate conservation strategies based upon these needs. These efforts began with the ISC’s Conservation Strategy (Thomas et al. 1990); they continued with the designation of critical habitat (USDI FWS 1992b), the Draft Recovery Plan (USDI FWS 1992a), and the Scientific Analysis Team report (Thomas et al. 1993), report of the Forest Ecosystem Management Assessment Team (Thomas and Raphael 1993); and they culminated with the NWFP (USDA FS/USDI BLM 1994a, b). Each conservation strategy was based upon the reserve design principles first articulated in the ISC’s report, which are summarized as follows.

 Species that are well distributed across their range are less prone to extinction than species confined to small portions of their range.  Large blocks of habitat, containing multiple pairs of the species, are superior to small blocks of habitat with only one to a few pairs.  Blocks of habitat that are close together are better than blocks far apart.  Habitat that occurs in contiguous blocks is better than habitat that is more fragmented.  Habitat between blocks is more effective as dispersal habitat if it resembles suitable habitat.

Federal Contribution to Recovery Since it was signed on April 13, 1994, the NWFP has guided the management of Federal forest lands within the range of the spotted owl (USDA FS and USDI BLM 1994a, 1994b). The NWFP was designed to protect large blocks of old growth forest and provide habitat for species that depend on those forests including the spotted owl, as well as to produce a predictable and sustainable level of timber sales. The NWFP included land use allocations which would provide for population clusters of spotted owls (i.e., demographic support) and maintain connectivity between population clusters. Certain land use allocations in the plan contribute to supporting population clusters: LSRs, Managed Late-successional 19

Areas, and Congressionally Reserved areas. Riparian Reserves, Adaptive Management Areas and Administratively Withdrawn areas can provide both demographic support and connectivity/dispersal between the larger blocks, but were not necessarily designed for that purpose. Matrix areas were to support timber production while also retaining biological legacy components important to old-growth obligate species (in 100-acre owl cores, 15 percent late-successional provision, etc. (USDA FS/USDI BLM 1994a, USDI FWS 1994b) which would persist into future managed timber stands.

The NWFP with its rangewide system of LSRs was based on work completed by three previous studies (Thomas et. al. 2006, pp. 279-280): the 1990 Interagency Scientific Committee (ISC) Report (Thomas et. al. 1990), the 1991 report for the Conservation of Late-successional Forests and Aquatic Ecosystems (Johnson et. al. 1991), and the 1993 report of the Scientific Assessment Team (Thomas et. al. 1993). In addition, the 1992 Draft Recovery Plan for the Northern Spotted Owl (USDI FWS 1992a) was based on the ISC report.

The Forest Ecosystem Management Assessment Team predicted, based on expert opinion, the spotted owl population would decline in the Matrix land use allocation over time, while the population would stabilize and eventually increase within LSRs as habitat conditions improved over the next 50 to 100 years (Thomas and Raphael 1993, p. II-31, USDA FS/USDI BLM 1994b, pp. 3&4-229). Based on the results of the first decade of monitoring, Lint et al.(2005, p. 18) could not determine whether implementation of the NWFP would reverse the spotted owl’s declining population trend because not enough time had passed to provide the necessary measure of certainty. Lint et al. (2005) concluded the results from the first decade of monitoring did not provide any reason to depart from the objective of habitat maintenance and restoration as described in the NWFP (Lint et al. 2005, p. 18; Noon and Blakesley 2006, p. 288). Bigley and Franklin (2004, pp. 6-34) suggested that more fuels treatments are needed in east-side forests to preclude large-scale losses of habitat to stand-replacing wildfires. Other stressors that occur in suitable habitat, such as the range expansion of the barred owl (already in action) and infection

In fact, information collected during the first fifteen years of the NWFP, affirms overall NRF habitat declines have been less than anticipated on federal lands (3.4% over fifteen years versus the 2.5 % per decade anticipated), and that although protecting habitat is still very important to the survival and recovery of the spotted owl the rapidly increasing barred owl population(s) are proving a confounding influence (Davis et al.2011). Recent reports about the status of the spotted owl offer few management recommendations to deal with these emerging threats. The arrangement, distribution, and resilience of the NWFP land use allocation system may prove to be the most appropriate strategy in responding to these unexpected challenges (Bigley and Franklin 2004, pp. 6-34). The revised recovery plan builds on the NWFP and recommends continued implementation of the NWFP and its standards and guidelines (USDI FWS 2011a, p I-1).

Under the NWFP, the agencies anticipated a decline of spotted owl populations during the first decade of implementation. Recent reports (Anthony et al. 2006, pp. 33-34) identified greater than expected spotted owl declines in Washington and northern portions of Oregon, and more stationary populations in southern Oregon and northern California. The reports did not find a direct correlation between habitat conditions and changes in vital rates of spotted owls at the meta-population scale. However, at the territory scale, there is evidence of negative effects to spotted owl fitness due to reduced habitat quantity and quality. Also, there is no evidence to suggest that dispersal habitat is currently limiting (Courtney et al. 2004, p. 9-12, Davis et al.2011). Even with the population decline, Courtney et al. (2004,

20 p. 9-15) noted that there is little reason to doubt the effectiveness of the core principles underpinning the NWFP conservation strategy.

The current scientific information, including information showing spotted owl population declines, indicates that the spotted owl continues to meet the definition of a (USDI FWS 2004a, p. 54). That is, populations are still relatively numerous over most of its historic range, which suggests that the threat of extinction is not imminent, and that the subspecies is not endangered; even though, in the northern part of its range population trend estimates are showing a decline.

In June, 2011, the Service published the Revised Recovery Plan for the Northern Spotted Owl (USDI FWS 2011a). The revised recovery plan determined that the 12 existing physiographic provinces meet the criteria of recovery units (USDI FWS 2011a, p. III 1-2). Recovery Criteria described in the revised recovery plan (USDI 2011a p. II-3) are measurable and achievable goals to be met via implementation of the recovery actions. Achieving the recovery criteria will take time and progress is intended to be measured over the life of the plan, not on a short-term basis. The recovery criteria are the same for all twelve identified recovery units: 1) stable population trend; 2) adequate population distribution; 3) continued maintenance and recruitment of spotted owl habitat, and; 4) post-delisting monitoring (USDI FWS 2011a p III-3).

The revised recovery plan identifies competition with barred owls, ongoing loss of spotted owl habitat as a result of timber harvest, loss or modification of habitat from uncharacteristic wildfire, and loss of amount and distribution of spotted owl habitat as a result of past activities and disturbances as the most pressing threats to the spotted owl (USDI FWS 2011a, p. II-2). To address these threats, the current recovery strategy includes five basic steps: 1. Development of a range-wide habitat modeling framework; 2. Barred owl management; 3. Monitoring and research; 4. Adaptive management; and 5. Habitat conservation and active forest restoration (USDI FWS 2011a, p. II-2).

The revised recovery plan lists recovery actions that address all five steps of the recovery strategy and further states that a habitat conservation network designed using the best available science is necessary to recover the spotted owl. The revised recovery plan indicates that the NWFP reserve network, in addition to other habitat conservation recommendations in the revised recovery plan (e.g., Recovery Actions 10, 32 and 6), meet that need in the near term (USDI FWS 2011a, p. II-3). Acknowledging that barred owls have reduced spotted owl site occupancy, reproduction, and survival, and that addressing this threat depends on initiating action as soon as possible, the revised recovery plan describes recovery actions which address research involving the competition between spotted and barred owls, experimental control of barred owls and, if recommended by research, management of barred owls (USDI FWS 2011a, p. II-4).

The final recovery plan (USDI FWS 2008a) recommended implementing a robust monitoring and research program for the spotted owl. The revised recovery plan encourages continuation of these efforts, laying out the following primary elements to evaluate progress toward meeting recovery criteria: monitoring spotted owl population trends, comprehensive barred owl research and monitoring, continued habitat monitoring; inventory of spotted owl distribution, and; explicit consideration for climate change mitigation goals consistent with recovery actions (USDI FWS 2011a, p. II-5). The revised recovery plan also strongly encourages land managers to aggressively implement recovery actions. In

21 other words, land managers should not be so conservative that, to avoid risk, they forego actions that are necessary to conserve the forest ecosystems that are necessary to the long-term conservation of the spotted owl. But they should also not be so aggressive that they subject spotted owls and their habitat to treatments where the long-term benefits do not clearly outweigh the short-term risks. Finding the appropriate balance to this dichotomy will remain an ongoing challenge for all who are engaged in spotted owl conservation (USDI FWS 2011a, p. II-12). The revised recovery plan estimates that recovery of the spotted owl could be achieved in approximately 30 years (USDI FWS 2011a, p. II-3).

Conservation Efforts on Non-Federal Lands In the report from the Interagency Scientific Committee (Thomas et al. 1990, p. 3), the draft recovery plan (USDI FWS 1992a, p. 272), and the report from the Forest Ecosystem Management Assessment Team (Thomas and Raphael 1993, pp. IV-189), it was noted that limited Federal ownership in some areas constrained the ability to form a network of old-forest reserves to meet the conservation needs of the spotted owl. In these areas in particular, non-Federal lands would be important to the range-wide goal of achieving conservation and recovery of the spotted owl. The Service’s primary expectations for private lands are for their contributions to demographic support (pair or cluster protection) to Federal lands, or their connectivity with Federal lands. In addition, timber harvest within each state is governed by rules that provide protection of spotted owls or their habitat to varying degrees. There are 17 current or ongoing Conservation Plans (CPs) including Habitat Conservation Plans (HCPs) and Safe Harbor Agreements (SHAs) that have incidental take permits issued for spotted owls—eight in Washington, three in Oregon, and six in California (USDI FWS 2011a, p. A-15). The CPs range in size from 76 acres to more than 1.8 million acres, although not all acres are included in the mitigation for spotted owls. In total, the HCPs cover approximately 3 million acres (9.4 percent) of the 32 million acres of non- Federal forest lands in the range of the spotted owl. The period of time that the HCPs will be in place ranges from 5 to 100 years; however, most of the HCPs are of fairly long duration. While each HCP is unique, there are several general approaches to mitigation of incidental take: Reserves of various sizes, some associated with adjacent Federal reserves Forest harvest that maintains or develops suitable habitat Forest management that maintains or develops dispersal habitat Deferral of harvest near specific sites

Washington. In 1996, the State Forest Practices Board adopted rules (Washington Forest Practices Board 1996) that would contribute to conserving the spotted owl and its habitat on non-Federal lands. Adoption of the rules was based in part on recommendations from a Science Advisory Group that identified important non-Federal lands and recommended roles for those lands in spotted owl conservation (Buchanan et al. 1994, p. ii). The 1996 rule package was developed by a stakeholder policy group and then reviewed and approved by the Forest Practices Board (Buchanan and Swedeen 2005, p. 9). Spotted owl-related HCPs in Washington generally were intended to provide demographic or connectivity support (USDI FWS 1992a, p. 272). There are over 2.1 million acres of land in six HCPs and two SHAs in Washington (USDI FWS 2011a, p. A- 15). Some of these conservation plans focus on providing nesting/roosting habitat throughout the area or in strategic locations, while others focus on providing connectivity through foraging and/or dispersal habitat. In addition, there is a long-term habitat management agreement covering 13,000 acres in which authorization of take was provided through an incidental take statement associated with a Federal land exchange (USDI FWS 2011a, p. A-15).

Oregon. The Oregon Forest Practices Act provides for protection of 70-acre core areas around sites occupied by an adult pair of spotted owls capable of breeding (as determined by recent protocol

22 surveys), but it does not provide for protection of spotted owl habitat beyond these areas (Oregon Department of Forestry 2007, p. 64). In general, no large-scale spotted owl habitat protection strategy or mechanism currently exists for non-Federal lands in Oregon. The three spotted owl-related HCPs currently in effect cover more than 300,000 acres of non-Federal lands. These HCPs are intended to provide some nesting habitat and connectivity over the next few decades (USDI FWS 2011a, p. A-16). On July 27, 2010, the Service completed a programmatic SHA with the Oregon Department of Forestry that will enroll up to 50,000 acres of non-federal lands within the State over 50 years. The primary intent of this programmatic SHA is to increase time between harvests and to lightly to moderately thin younger forest stands that are currently not habitat to increase tree diameter and stand diversity (USDI FWS 2011a p. A-16).

California. The California State Forest Practice Rules, which govern timber harvest on private lands, require surveys for spotted owls in suitable habitat and to provide protection around activity centers (California Department of Forestry and Fire Protection 2007, pp. 85-87). Under the Forest Practice Rules, no timber harvest plan can be approved if it is likely to result in incidental take of federally listed species, unless the take is authorized by a Federal incidental take permit (California Department of Forestry and Fire Protection 2007, pp. 85-87). The California Department of Fish and Game initially reviewed all timber harvest plans to ensure that take was not likely to occur; the Service took over that review function in 2000. Several large industrial owners operate under spotted owl management plans that have been reviewed by the Service and that specify basic measures for spotted owl protection. Four HCPs and two SHAs authorizing take of spotted owls have been approved; these HCPs cover more than 622,000 acres of non-Federal lands. Implementation of these plans is intended to provide for spotted owl demographic and connectivity support to NWFP lands (USDI FWS 2011a, p. A-16).

Current Condition of the Spotted Owl

The current condition of the species incorporates the effects of all past human activities and natural events that led to the present-day status of the species and its habitat (USDI FWS/USDC NMFS 1998, pp. 4-19).

Range-wide Habitat and Population Trends

Habitat Baseline The 1992 Draft Spotted Owl Recovery Plan estimated approximately 8.3 million acres of spotted owl habitat remained range-wide (USDI FWS 1992a, p. 37). However, reliable habitat baseline information for non-Federal lands is not available (Courtney et al. 2004, p. 6-5). The Service has used information provided by the Forest Service, Bureau of Land Management, and to update the habitat baseline conditions on Federal lands for spotted owls on several occasions since the spotted owl was listed in 1990. The estimate of 7.4 million acres used for the NWFP in 1994 (USDA FS/USDI BLM 1994b, p. G-34) was believed to be representative of the general amount of spotted owl habitat on these lands. This baseline has been used to track relative changes over time in subsequent analyses, including those presented here.

In 2010, a new map depicting suitable spotted owl habitat throughout the range of the spotted owl was produced as a result of the NWFP’s effectiveness monitoring program (Davis et al.2011). However, the spatial resolution of this new habitat map currently makes it unsuitable for tracking habitat effects at the scale of individual projects. The Service is evaluating the map for future use in tracking habitat trends. Additionally, there continues to be no reliable estimates of spotted owl habitat on non-Federal

23 lands; consequently, consulted-on acres can be tracked, but not evaluated in the context of change with respect to a reference condition on non-Federal lands. The production of the monitoring program habitat map does, however, provide an opportunity for future evaluations of trends in non-Federal habitat.

NWFP Lands Analysis 1994 – 2001 In 2001, the Service conducted an assessment of habitat baseline conditions, the first since implementation of the NWFP (USDI FWS 2001, p. 1). This range-wide evaluation of habitat, compared to the FSEIS, was necessary to determine if the rate of potential change to spotted owl habitat was consistent with the change anticipated in the NWFP. In particular, the Service considered habitat effects that were documented through the section 7 consultation process since 1994. In general, the analytical framework of these consultations focused on the reserve and connectivity goals established by the NWFP land-use allocations (USDA FS/ USDI BLM 1994a, p. 6), with effects expressed in terms of changes in suitable spotted owl habitat within those land-use allocations. The Service determined that actions and effects were consistent with the expectations for implementation of the NWFP from 1994 to June, 2001 (USDI FWS 2001, p. 32).

Range-wide Analysis 1994 – January 29, 2013 This section updates the information considered in USDI FWS (2001), relying particularly on information in documents the Service produced pursuant to section 7 of the Act and information provided by NWFP agencies on habitat loss resulting from natural events (e.g., fires, wind throw, insect and disease). To track impacts to spotted owl habitat, the Service designed the Consulted on Effects Database which records impacts to spotted owls and their habitat at a variety of spatial and temporal scales. Data are entered into the Consulted on Effects Database under various categories including, land management agency, land-use allocation, physiographic province, and type of habitat affected.

In 1994, about 7.4 million acres of suitable spotted owl habitat were estimated to exist on Federal lands managed under the NWFP. As of January 29, 2013, the Service had consulted on the proposed removal/downgrading of approximately 194,305 acres (Table 1) or nearly 2.7 percent of 7.4 million acres (Table 2) of spotted owl NRF habitat on Federal lands. Of the total Federal acres consulted on for removal/downgrading, approximately 194,305 acres or 2.62 percent of 7.4 million acres of spotted owl NRF habitat were removed/downgraded as a result of timber harvest (Table 1). These changes in suitable spotted owl habitat are consistent with the expectations for implementation of the NWFP (USDA FS/USDI BLM 1994a).

Habitat loss from Federal lands due to management activities has varied among the individual provinces with most of the impacts concentrated within the Non-Reserve relative to the Reserve land-use allocations (Table 2). When habitat loss is evaluated as a proportion of the affected acres range-wide, the most pronounced losses have occurred within Oregon (72%), especially within its Klamath Mountains (37%) and Cascades (East and West) (35%) Provinces (Table 2), followed by much smaller habitat losses in Washington (9%) and California (18%) (Table 1). When habitat loss is evaluated as a proportion of provincial baselines, the Oregon Klamath Mountains (20.62%), Oregon Cascades East (13%), and the California Klamath (6.92%) all have proportional losses greater than the average loss of habitat across all provinces (5.95%) (Table 2).

From 1994 through January 29, 2013, habitat lost due to natural events was estimated at approximately 246,111 acres range-wide (Table 2). About two-thirds of this loss was attributed to the Biscuit Fire that burned over 500,000 acres in southwest Oregon (Rogue River basin) and northern California in 2002.

24

This fire resulted in a loss of approximately 100,000 acres of spotted owl habitat (Table 2), including habitat within five LSRs. Approximately 18,630 acres of spotted owl habitat were lost due to the B&B Complex and Davis Fires in the East Cascades Province of Oregon (Table 2).

Because there is no comprehensive spotted owl habitat baseline for non-NWFP Federal lands and non- Federal lands, there is little available information regarding spotted owl habitat trends on these lands. Yet, we do know that Service consultations conducted since 1992, have documented the eventual loss of 472,772 acres (Table 3) acres of habitat on non-NWFP Federal lands and non-Federal lands. Approximately 63 percent of these losses have yet to be realized because they are part of large-scale, long-term HCPs/SHAs. Combining effects on Federal and non-Federal lands, the Service had consulted on the proposed removal/downgrading of approximately 667,077 acres of spotted owl habitat range- wide, resulting from all management activities, as of January 29, 2013 (Table 2).

Table 1. Range-wide Aggregate of Changes to NRF1 Habitat Acres from Activities Subject to Section 7 Consultations and Other Causes. LAND OWNERSHIP CONSULTED ON HABITAT CHANGES2 OTHER HABITAT CHANGES3 Removed/ Removed/ Downgraded Maintained Downgraded Maintained NWFP (FS,BLM,NPS) 194,305 535,109 246,111 39,720 Bureau of Indian Affairs / Tribes 108,210 28,372 2,398 0 Habitat Conservation Plans/Safe Harbor Agreements 295,889 14,430 N/A N/A Other Federal, State, County, Private Lands 68,673 5,620 279 0 Total Changes 667,077 605,425 248,788 39,720 1. Nesting, roosting, foraging (NRF) habitat. In California, suitable habitat is divided into two components; nesting - roosting (NR) habitat, and foraging (F) habitat. The NR component most closely resembles NRF habitat in Oregon and Washington. Due to differences in reporting methods, effects to suitable habitat compiled in this, and all subsequent tables include effects for nesting, roosting, and foraging (NRF) for 1994-6/26/2001. After 6/26/2001 suitable habitat includes NRF for Washington and Oregon but only nesting and roosting (NR) for California. 2. Includes both effects reported in USDI FWS 2001 and subsequent effects reported in the Spotted Owl Consultation Effects Tracking System (web application and database.) 3. Includes effects to suitable NRF habitat (as generally documented through technical assistance, etc.) resulting from wildfires (not from suppression efforts), insect and disease outbreaks, and other natural causes, private timber harvest, and land exchanges not associated with consultation.

25

Table 2. Aggregate Results of All Adjusted, Suitable Habitat (NRF1) Acres Affected by Section 7 Consultation on NWFP Lands for the spotted owl; Baseline and Summary of Effects by State, Physiographic Province and Land Use Function. PHYSIOGRAPHIC HABITAT REMOVED/DOWNGRADED4 2 EVALUATION PROVINCE 3 BASELINE Land Use Allocations Habitat Loss to % PROVINCIAL % RANGE- Non- Natural BASELINE WIDE Total Reserves5 Reserves6 Total Events7 Total AFFECTED EFFECTS W Eastern 706,849 4,557 6,392 10,949 14,307 25,256 3.57 5.73 A Cascades Olympic 560,217 869 1,711 2,580 299 2,879 0.51 0.65 Peninsula Western 1,112,480 778 11,579 12,357 3 12,360 1.11 2.81 Cascades OR Cascades 443,659 2,589 14,368 16,957 40,884 57,841 13.04 13.13 East Cascades 2,046,472 3,872 66,501 70,373 24,583 94,956 4.64 21.56 West Coast Range 516,577 447 4,002 4,449 66 4,515 0.87 1.03 Klamath 785,589 3,923 56,360 60,283 101,676 161,959 20.62 36.77 Mountains Willamette 5,658 0 0 0 0 0 0 0 Valley CA Cascades 88,237 10 4,820 4,830 329 5,159 5.85 1.17 Coast 51,494 464 79 543 275 818 1.59 0.19 Klamath 1,079,866 1,546 9,438 10,984 63,689 74,673 6.92 16.96 Total 7,397,098 19,055 175,250 194,305 246,111 440,416 5.95 100 1. Nesting, roosting, foraging (NRF) habitat. In California, suitable habitat is divided into two components; nesting - roosting (NR) habitat, and foraging (F) habitat. The NR component most closely resembles NRF habitat in Oregon and Washington. Due to differences in reporting methods, effects to suitable habitat compiled in this, and all subsequent tables include effects for nesting, roosting, and foraging (NRF) for 1994-6/26/2001. After 6/26/2001 suitable habitat includes NRF for Washington and Oregon but only nesting and roosting (NR) for California. 2. Defined by the Northwest Forest Plan as the twelve physiographic provinces, as presented in Figure 3&4-1 on page 3&4-16 of the FSEIS. The WA Western Lowlands and OR Willamette Valley provinces are not listed as they are not expected to contribute to recovery. 3. 1994 FSEIS baseline (USDA FS/USDI BLM 1994b). 4. Includes both effects reported in USDI FWS 2001 and subsequent effects reported in the Northern Spotted owl Consultation Effects Tracking System (web application and database.) 5. Land-use allocations intended to provide large blocks of habitat to support clusters of breeding pairs. (LSR, MLSA, CRA) 6. Land-use allocations intended to provide habitat to support movement of spotted owls among reserves. (AWA, AMA, MX) 7. Acres for all physiographic provinces, except the Oregon Klamath Mountains, are from the Scientific Evaluation of the Status of the Northern Spotted Owl (Courtney et al. 2004) and subsequent effects entered into the Northern Spotted Owl Consultation Effects Tracking System. Acres for the Oregon Klamath Mountains province are from the biological assessment entitled: Fiscal year 2006-2008 programmatic consultation: re-initiation on activities that may affect listed species in the Rogue-River/South Coast Basin, Medford BLM, and Rogue-Siskiyou National Forest and from subsequent effects entered into the Northern Spotted Owl Consultation Effects Tracking System.

26

Other Habitat Trend Assessments In 2005, the Washington Department of Wildlife released the report, “An Assessment of Spotted Owl Habitat on Non-Federal Lands in Washington between 1996 and 2004” (Pierce et al. 2005). This study estimates the amount of spotted owl habitat in 2004 on lands affected by state and private forest practices. The study area is a subset of the total Washington forest practice lands, and statistically- based estimates of existing habitat and habitat loss due to fire and timber harvest are provided. In the 3.2-million acre study area, Pierce et al. (2005, p. 88) estimated there was 816,000 acres of suitable spotted owl habitat in 2004, or about 25 percent of their study area. Based on their results, Pierce and others (2005, p. 98) estimated there were less than 2.8 million acres of spotted owl habitat in Washington on all ownerships in 2004. Most of the suitable owl habitat in 2004 (56%) occurred on Federal lands, and lesser amounts were present on state-local lands (21%), private lands (22%) and tribal lands (1%). Most of the harvested spotted owl habitat was on private (77%) and state-local (15%) lands. A total of 172,000 acres of timber harvest occurred in the 3.2 million-acre study area, including harvest of 56,400 acres of suitable spotted owl habitat. This represented a loss of about 6 percent of the owl habitat in the study area distributed across all ownerships (Pierce et al. 2005, p. 91). Approximately 77 percent of the harvested habitat occurred on private lands and about 15 percent occurred on State lands. Pierce and others (2005, p. 80) also evaluated suitable habitat levels in 450 spotted owl management circles (based on the provincial annual median spotted owl home range). Across their study area, they found that owl circles averaged about 26 percent suitable habitat in the circle across all landscapes. Values in the study ranged from an average of 7 percent in southwest Washington to an average of 31 percent in the east Cascades, suggesting that many owl territories in Washington are significantly below the 40 percent suitable habitat threshold used by the State as a viability indicator for spotted owl territories (Pierce et al. 2005, p. 90).

Moeur et al. 2005 (p. 110) estimated an increase of approximately 1.25 to 1.5 million acres of medium and large older forest (greater than 20 inches dbh, single and multi-storied canopies) on Federal lands in the Northwest Forest Plan (NWFP) area between 1994 and 2003. The increase occurred primarily in the lower end of the diameter range for older forest. The net area in the greater than 30 inch dbh size class increased by only an estimated 102,000 to 127,000 acres (Moeur et al. 2005, p. 100). The estimates were based on change-detection layers for losses due to harvest and fire and remeasured inventory plot data for increases due to ingrowth. Transition into and out of medium and large older forest over the 10-year period was extrapolated from inventory plot data on a subpopulation of Forest Service land types and applied to all Federal lands. Because size class and general canopy layer descriptions do not necessarily account for the complex forest structure often associated with spotted owl habitat, the significance of these acres to spotted owl conservation remains unknown.

In 2011, Davis et al. produced the second in a series of monitoring reports on northern spotted owl population and habitat trends on Northwest Forest Plan administered lands. They summarized demographic analyses from Forsman et al. (2011) discussed below under trends in numbers, distribution and reproduction, and reported on a new effort using remotely sensed data from 1994 to 2007 to develop “habitat suitability” models, and ultimately suitable habitat maps for the entire range of the northern spotted owl for each of these time periods. They also created change-detection maps and reported on the cause of habitat change during this time period. The authors suggest that because of improvements in remotely sensed vegetation, and change-detection mapping, their habitat maps represent the best available information and should replace the baseline versions used for the first monitoring report. Davis et al. (2011) estimated 8.9 million acres of suitable habitat for the 1994 baseline map, as compared to 7.4 million acres estimated by FEMAT in 1994, and 10.3 million acres estimated by Lint (2005) for the 10-year report.

27

Davis et al. (2011) were not able to report on gains in nesting/roosting habitat suitability due to issues with current technology, and the need for additional time to capture the slow process of forest succession. However, they were able to report on gains in recruitment of younger forests or dispersal habitat. They estimated a gain of about 1.26 million ac of dispersal habitat, with the greatest increases in non-reserves than reserves. The largest increase in dispersal habitat was in the Oregon Coast Range province.

Davis et al. (2011) estimated that nesting/roosting habitat declined by 3.4 percent (298,600 ac) rangewide on federal lands since 1994, which is less than the anticipated rate of habitat loss under the NWFP of 2.5 percent per decade. Most of the loss (79 percent) occurred within reserves and was the result of wildfires. Wildfires also were responsible for about half of the loss in non-reserves. Timber harvest accounted for about 45 percent (37,400 ac) in non-reserves, and 7 percent (16,000 ac) in reserves. The Oregon Klamath province lost the most nesting/roosting habitat (93,730 ac) due to the Biscuit Fire in 2002. They estimated a rangewide loss of about 417,000 ac of dispersal habitat, but like nesting/roosting habitat, most of the loss of dispersal habitat was due to wildfire.

Davis et al. (2011) also created a wildfire suitability (likelihood) map for large fires throughout the range of the northern spotted owl. Their goal was to identify landscape-scale areas where large wildfires are more probable. They report that the California Klamath province has the most owl habitat in fire-prone landscapes, followed by the Oregon Western Cascades and Oregon Klamath provinces.

Spotted Owl Numbers, Distribution, and Reproduction Trends There are no estimates of the size of the spotted owl population prior to settlement by Europeans. Spotted owls are believed to have inhabited most old-growth forests or stands throughout the Pacific Northwest, including northwestern California, prior to beginning of modern settlement in the mid-1800s (USDI FWS 1989, pp. 2-17). According to the final rule listing the spotted owl as threatened (USDI FWS 1990, p. 26118), approximately 90 percent of the roughly 2,000 known spotted owl breeding pairs were located on Federally managed lands, 1.4 percent on State lands, and 6.2 percent on private lands; the percent of spotted owls on private lands in northern California was slightly higher (USDI FWS 1989, pp. 4-11; Thomas et al. 1990, p. 64).

The current range of the spotted owl extends from southwest British Columbia through the Cascade Mountains, coastal ranges, and intervening forested lands in Washington, Oregon, and California, as far south as Marin County (USDI FWS 1990, p. 26115). The range of the spotted owl is partitioned into 12 physiographic provinces (Figure 2) based on recognized landscape subdivisions exhibiting different physical and environmental features (USDI FWS 1992a, p. 31). The spotted owl has become rare in certain areas, such as British Columbia, southwestern Washington, and the northern coastal ranges of Oregon.

As of July 1, 1994, there were 5,431 known site-centers of spotted owl pairs or resident singles: 851 sites (16 percent) in Washington, 2,893 sites (53 percent) in Oregon, and 1,687 sites (31 percent) in California (USDI FWS 1995, p. 9495). By June 2004, the number of territorial spotted owl sites in Washington recognized by the Washington Department of Fish and Wildlife was 1,044 (Buchanan and Swedeen 2005, p. 37). The actual number of currently occupied spotted owl locations across the range is unknown because many areas remain unsurveyed (USDI FWS 2008a, p. 44). In addition, many historical sites are no longer occupied because spotted owls have been displaced by barred owls, timber harvest, or severe fires, and it is possible that some new sites have been established due to reduced timber

28 harvest on Federal lands since 1994. The totals in USDI FWS (1995, p. 9495) represent the cumulative number of locations recorded in the three states, not population estimates.

Because the existing survey coverage and effort are insufficient to produce reliable range-wide estimates of population size, demographic data are used to evaluate trends in spotted owl populations. Analysis of demographic data can provide an estimate of the finite rate of population change (λ) (lambda), which provides information on the direction and magnitude of population change. A λ of 1.0 indicates a stationary population, meaning the population is neither increasing nor decreasing. A λ of less than 1.0 indicates a decreasing population, and a λ of greater than 1.0 indicates a growing population. Demographic data, derived from studies initiated as early as 1985, have been analyzed periodically (Anderson and Burnham 1992; Burnham et al. 1994: Forsman et al. 1996; Anthony et al. 2006 and Forsman et al. 2011) to estimate trends in the populations of the spotted owl.

In January 2009, two meta-analyses modeled rates of population change for up to 24 years using the re- parameterized Jolly-Seber method (λRJS). One meta-analysis modeled the 11 long-term study areas (Figure 3), while the other modeled the eight study areas that are part of the effectiveness monitoring program of the NWFP (Forsman et al. 2011).

Figure 3. Northern spotted owl demographic study areas (Forsman et al. 2011).

29

Table 3. Demographic results from spotted owl demographic study areas (adapted from Forsman et al. 2011). Apparent Population 1 2 Study Area Fecundity Survival λRJS change Cle Elum Declining Declining 0.937 Declining Rainier Increasing Declining 0.929 Declining Olympic Stable Declining 0.957 Declining Declining since Coast Ranges Increasing 1998 0.966 Declining Declining since HJ Andrews Increasing 1997 0.977 Declining Declining since Tyee Stable 2000 0.996 Stationary Klamath Declining Stable 0.990 Stationary Declining since Southern Cascades Declining 2000 0.982 Stationary NW California Declining Declining 0.983 Declining Declining since Hoopa Stable 2004 0.989 Stationary Green Diamond Declining Declining 0.972 Declining 1Apparent survival calculations are based on model average. 2 Population trends are based on estimates of realized population change, which differs slightly from λRJS.

Point estimates of λRJS were all below 1.0 and ranged from 0.929 to 0.996 for the 11 long-term study areas (Table 3). There was strong evidence that populations declined on 7 of the 11 areas (Forsman et al. 2011), these areas included Rainier, Olympic, Cle Elum, Coast Range, HJ Andrews, Northwest California and Green Diamond. On other four areas (Tyee, Klamath, Southern Cascades, and Hoopa), populations were either stable, or the precision of the estimates was not sufficient to detect declines.

The weighted mean λRJS for all of the 11 study areas was 0.971 (standard error [SE] = 0.007, 95 percent confidence interval [CI] = 0.960 to 0.983), which indicated an average population decline of 2.9 percent per year from 1985 to 2006. This is a lower rate of decline than the 3.7 percent reported by Anthony et al. (2006), but the rates are not directly comparable because Anthony et al. (2006) examined a different series of years and because two of the study areas in their analysis were discontinued and not included in Forsman et al. (2011). Forsman et al. (2011) explains that the indication populations were declining was based on the fact that the 95 percent confidence intervals around the estimate of mean lambda did not overlap 1.0 (stable) or barely included 1.0.

The mean λRJS for the eight demographic monitoring areas (Cle Elum, Olympic, Coast Range, HJ Andrews, Tyee, Klamath, Southern Cascades and Northwest California) that are part of the effectiveness monitoring program of the NWFP was 0.972 (SE = 0.006, 95 percent CI = 0.958 to 0.985), which indicated an estimated decline of 2.8 percent per year on Federal lands with the range of the spotted owl. The weighted mean estimate λRJS for the other three study areas (Rainier, Hoopa and Green Diamond) was 0.969 (SE = 0.016, 95 percent CI = 0.938 to 1.000), yielding an estimated average decline of 3.1 percent per year. These data suggest that demographic rates for spotted owl populations on Federal lands were

30 somewhat better than elsewhere; however, this comparison is confounded by the interspersion of non- Federal land in study areas and the likelihood that spotted owls use habitat on multiple ownerships in some demography study areas.

The number of populations that declined and the rate at which they have declined are noteworthy, particularly the precipitous declines in the Olympic, Cle Elum, and Rainier study areas in Washington and the Coast Range study area in Oregon. Estimates of population declines in these areas ranged from 40 to 60 percent during the study period through 2006 (Forsman et al. 2011). Spotted owl populations on the HJ Andrews, Northwest California, and Green Diamond study areas declined by 20-30 percent whereas the Tyee, Klamath, Southern Cascades, and Hoopa study areas showed declines of 5 to 15 percent.

Decreases in adult apparent survival rates were an important factor contributing to decreasing population trends. Forsman et al. (2011) found apparent survival rates were declining on 10 of the study area with the Klamath study area in Oregon being the exception. Estimated declines in adult survival were most precipitous in Washington where apparent survival rates were less than 80 percent in recent years, a rate that may not allow for sustainable populations (Forsman et al. 2011). In addition, declines in adult survival for study areas in Oregon have occurred predominately within the last five years and were not observed in the previous analysis by Anthony et al. 2006. Forsman et al. (2011) express concerns by the collective declines in adult survival across the subspecies range because spotted owl populations are most sensitive to changes in adult survival.

There are few spotted owls remaining in British Columbia. Chutter et al. (2004, p. v) suggested immediate action was required to improve the likelihood of recovering the spotted owl population in British Columbia. So, in 2007, personnel in British Columbia captured and brought into captivity the remaining 16 known wild spotted owls (USDI FWS 2008b, p. 48). Prior to initiating the captive-breeding program, the population of spotted owls in Canada was declining by as much as 10.4 percent per year (Chutter et al. 2004, p. v). The amount of previous interaction between spotted owls in Canada and the United States is unknown.

Physiographic Provinces The Action Area lies entirely within the Oregon Cascades West Physiographic Province (OCWP), which extends from the California-Oregon border to the Oregon-Washington border. The OCWP is located in the geographic center of the spotted owl’s range and in addition to the interstate links on it north and south borders, it provides links to the Oregon Coast Ranges, Oregon Cascades East and the Oregon Klamath Mountains Physiographic Provinces. The 1994 FSEIS baseline (USDA FS/USDI BLM 1994b) was 2,046,472 acres of spotted owl NRF habitat on federal lands within the OCWP, although consultation has been completed on the removal or downgrading of 70,373 acres through January 29, 2013, which, in addition to the loss of approximately 24,583 acres of NRF habitat due to natural events, has resulted in the decrease of 4.64 percent (94,956 acres) of the provincial baseline (Table 2).

Northwest Forest Plan (NWFP) The Forest is located within lands managed under the NWFP, which established the current conservation strategy for the spotted owl on federal lands. It is designed around the conservation needs of the spotted owl and based on the designation of a variety of land-use allocations whose objectives are either to provide for population clusters (i.e., demographic support) or to maintain connectivity between population clusters. Several land-use allocations are intended to contribute primarily to supporting population clusters: LSRs, Managed Late-Successional Areas (MLSAs), Congressionally

31

Reserved Areas (CRAs), Managed Pair Areas and Reserve Pair Areas. The remaining LUAs [Matrix, AMAs, Riparian Reserves (RRs), Connectivity Blocks, and AWAs] provide connectivity between habitat blocks intended for demographic support.

The range-wide system of LSRs set up under the NWFP captures the variety of ecological conditions within the 12 different physiographic provinces to which spotted owls are adapted. This design reduces the potential for extinction due to large catastrophic events in a single province. Multiple, large LSRs in each province reduce the potential that spotted owls will be extirpated in any individual province and reduce the potential that large wildfires or other events will eliminate all habitat within a LSR. In addition, LSRs are generally arranged and spaced so that spotted owls may disperse to two or more adjacent LSRs and/or Wilderness complexes. This network of reserves reduces the likelihood that catastrophic events will impact habitat connectivity and population dynamics within and between provinces.

It has been nineteen years since the adoption of the NWFP. Thomas et al. (1990) argued that the spotted owl population trend should stabilize at a lower equilibrium size sometime within the next 100 years. During the interim, there was an expectation the rate of decline would slowly decrease as habitat loss was arrested and new habitat regenerated in the habitat conservation areas. Similarly, the NWFP predicted a continuing decline of spotted owls until such time as new habitat developed (over a course of decades) (Appendix J of FSEIS) (Courtney et al. 2004). Lint et al. (2005) concluded that during the first ten years of the NWFP, the spotted owl habitat prognosis is seemingly correct. Anthony et al. (2006) stated spotted owl populations appeared to be stationary in several study areas as a result of high survival and stable fecundity rates, although federal study areas appear more stationary than those composed largely of private land. While the habitat provision of the NWFP is a necessary condition for spotted owls, it may not be a wholly sufficient provision (Courtney et al. 2004), given the spotted owl population declines observed in Washington (Anthony et al. 2006, Forsman et al.2011). Information collected during the first fifteen years of the NWFP, affirms overall NRF habitat declines have been less than anticipated on federal lands (3.4% over fifteen years versus the 2.5% per decade anticipated), and that although protecting habitat is still very important to the survival and recovery of the spotted owl the rapidly increasing barred owl population(s) are proving a confounding influence (Davis et al.2011).

Status within Spotted Owl Demography Study Areas

The Action Area is located between three long-term spotted owl demographic study areas (DSA), the Tyee, HJ Andrews and Southern Cascades, out of fourteen throughout the spotted owl’s range. We believe information from these areas provides inference regarding the condition of spotted owls within their respective provinces (Anthony et al. 2006) and likely within the Action Area. It is important to draw a distinction between conclusions drawn from range-wide demographic analyses, in particular Anthony et al. 2006 and Forsman et al. 2011, and the trends apparent in local data from individual DSAs. Whereas the range-wide estimates of population change (λ) (Table 3) are largely based on individual spotted owl survival probabilities, the annual reports from the individual DSAs focus on individual spotted owl nest site/territory occupancy trends. It is quite plausible that in declining populations of long-lived K-selected species, the high re-capture probabilities of aging individuals, and the population trend estimates derived from them, will somewhat mask, or lag behind site occupancy trends for pairs, which are much more responsive to year-to-year decreases (J. Reid 2011, pers. comm.).

Tyee Demography Study Area

32

The Tyee Study Area (TYE DSA) is located mostly on the western portion of the neighboring Roseburg District BLM, approximately 42 miles west of the Action Area. It was established to provide spotted owl demographic information for the Oregon Coast Ranges (Forsman et al. 2012). On this study area in 2011, pair occupancy decreased to 32 pairs, which is not only the lowest level observed during the 20+- year study, it is less than 50 percent of the number of pairs documented as recently as 2005. The total number of non-juvenile spotted owls on the study area was again the lowest ever observed: 91 individuals were documented (Forsman et al.2012). These changes in occupancy levels continue long- term trends apparent in the TYE DSA and may reflect a decrease in population size. In 2011 there were only three nests and only 2 young were documented (Forsman et al.2012).

Relative to the thirteen other spotted owl study areas represented in the 2009 meta-analysis, the demographic parameters estimated for the spotted owl population in the TYE DSA were at intermediate levels (Forsman et al. 2011). The greatest population declines were reported in the Oregon Coast Range and in all Washington study areas. These study areas had negative time trends in fecundity and/or survival. In contrast, spotted owl populations in southwestern Oregon and northern California appeared to be relatively stationary based on the 95 percent confidence intervals. However, as described previously, data collected since 1990 indicates downward trends in many areas, including all three of the study areas surrounding the Action Area. For example, the 91 spotted owls located within the TYE DSA in 2011 represents approximately 64 percent of the number of individuals documented in 1990 and was the lowest number of spotted owls detected since the study began. Barred owl detections (and likely numbers) continue to increase in the study area (Figure 4).

Figure 4. Spotted owl territories and reproduction and Barred owl territories relationship on the Tyee Density Study Area (Forsman et al. 2013). 160 Relationship of barred owls to spotted owl nesting success, 140 Tyee DSA 1990-2012

R² = 0.9446 120 Number of territories with STOC young produced

Average number of STOC territories producing 100 Number of territories with STVA

80 Expon. (Number of territories with STVA)

60 Numberofterritories

40

20

0 1990 1992 1994 1996 1998 2000 2002 2004 2006 2008 2010 2012

33

H.J. Andrews-Central Cascades Demography Study Area

The Central Cascades Study Area (HJA DSA) is centered on the H.J. Andrews Experimental Forest, on the McKenzie River Ranger District, Willamette National Forest (WNF), approximately 48 miles north- northeast of the Action Area. It was established to provide spotted owl demographic information for the Central west Oregon Cascades Mountains. Dugger et al.(2012 p.5) presented evidence that this spotted owl population continued to decline on the HJA; 2011 represented the lowest number of either a pair or resident single spotted owl per site surveyed since 1996 and the highest number of sites surveyed without spotted owl presence since the study began in 1987 (Figure 5). The proportion of surveyed sites with barred owl detections (47%) was greater in 2010 than in any previous year. The percentage of sites where at least one response from a barred owl was recorded decreased from 2010 (47%) to 2011 (43%) but remains higher than in any other year since the initiation of the study. In 2010, spotted owls reproduced on only one site out of 174 sites checked, and in 2011, forty-three spotted owl pairs and four resident single females were surveyed for reproductive status by 31 August 2011 – only one pair reproduced (2 young). However, up through the 2009 meta-analysis, fecundity was shown to be increasing (Table 3) (Forsman et al. 2011).

Figure 5. Percentage of sites occupied by pairs of northern spotted owls compared among land use allocations in the HJ Andrews-Central Cascades study area, Willamette National Forest, Oregon from 1997 – 2012 (Dugger et al. 2013).

South Cascades Demographic Study Area

In the South Cascades Study Area (60 miles south of the Action Area) spotted owls occupied approximately 65 percent of the 170 spotted owl sites surveyed in 2011 (Dugger et al. 2012). While the relative proportion of sites occupied between years has been similar, this is confounded by the increase

34 in the number of sites monitored. It should be noted, that overall, the percent occupancy rate of sites for this demographic area have been in substantial decline during the past decade. Concurrently, the fecundity rate (number of young produced per year) is also showing a negative trend and apparent survival has been in decline since 2000 (Forsman et al. 2011).

Late Successional Reserves (LSRs)

The occupancy and fecundity trends reported for the two surrounding DSAs are calculated differently in each case, making direct comparisons between DSAs problematic. However, there appear to be general downward trends in LSR contributions. For example, the proportion of surveyed sites occupied in the LSRs for the HJA DSA has been declining for several years where in 2010 only 39 percent, and in 2011 only 30 percent of the sites with pairs were in the LSRs (Dugger et al. 2012) (Figure 5).

2012 Revised Spotted Owl Critical Habitat The proposed action is planned within revised designated critical habitat for the spotted owl (USDI FWS 2012), specifically within critical habitat unit West Cascades South, subunit WCS #5 (WCS-5) (Table 4).

Table 4. Environmental Baseline for Revised Northern Spotted Owl Critical Habitat for the West Cascades South Unit, Sub-Unit WCS-5. Critical Habitat Sub-Unit Total federal Federal NRF habitat Federal Dispersal habitat Number acres acreage acreage WCS-5 356,718 219,053 (61%) 102,286 (29%)

Critical Habitat Unit West Cascades South The West Cascades South critical habitat unit contains 1,397,642 ac (565,608 ha) (Glenn pers. comm) and comprises six subunits. This unit consists of the southern portion of the Western Cascades Section M242B, based on section descriptions of forest types from Ecological Subregions of the United States (McNab and Avers 1994a, Section M242B), and extends from the Columbia River south to include the North Umpqua River watershed. This region was separated from the northern section due to its relatively milder temperatures, reduced summer precipitation due to the influence of the Willamette Valley to the west, lower elevations, and greater proportion of western hemlock/Douglas-fir forest. The southern portion of this region exhibits a gradient between Douglas-fir/ western hemlock and increasing Klamath-like vegetation (mixed conifer/evergreen hardwoods) which continues across the Umpqua divide area. The southern boundary of this region is novel and reflects a transition to mixed-conifer forest (Franklin and Dyrness 1988, pp. 23-24, 137-143). The importance of Douglas-fir dwarf mistletoe increases to the south in this region, but most spotted owl nest sites are in decadent large trees, and occasionally nests of other raptors.

Subunit WCS-5. The WCS-5 subunit consists of approximately 356,718 ac (144,359 ha) in Lane and Douglas Counties, Oregon, and comprises only Federal lands managed by the Forest Service under the NWFP (USDA FS/USDI BLM 1994a, entire, Figure 6). Congressionally reserved natural areas in Federal ownership are proposed for exclusion. Special management considerations or protection are required in this subunit to address threats from current and past timber harvest and competition with barred owls. This subunit is expected to function primarily for demographic support to the overall population, as well as north- south and east-west connectivity between subunits and CHUs. Our evaluation of sites known to be

35 occupied at the time of listing indicates that approximately 83 percent of the area of WCS-5 was covered by verified spotted owl home ranges at the time of listing. When combined with likely occupancy of suitable habitat and occupancy by non-territorial owls and dispersing subadults, we consider this subunit to have been largely occupied at the time of listing. In addition, there may be some smaller areas of younger forest within the habitat mosaic of this subunit that were unoccupied at the time of listing. We have determined that all of the unoccupied and likely occupied areas in this subunit are essential for the conservation of the species to meet the recovery criterion that calls for the continued maintenance and recruitment of spotted owl habitat (USDI FWS 2011a, p. ix). The increase and enhancement of spotted owl habitat is necessary to provide for viable populations of spotted owls over the long term by providing for population expansion, successful dispersal, and buffering from competition with the barred owl (Strix varia).

The Umpqua National Forest Baseline The Umpqua National Forest (NF) baseline habitat is described in terms of acres by land allocation in Table 5. The Umpqua NF is comprised of 986,135 acres of primarily mixed conifer forest. The majority of the Umpqua NF is capable habitat with the exception of the small amount of non-forested land (meadows, lakes, rivers, rocky outcrops, etc.) and upper elevation sub alpine fir and lodgepole pine forests which are not widely utilized by spotted owls (USDI 2008a). However, 53% of the Umpqua NF currently serves as suitable habitat of which 59% is within protected habitat. The majority of the capable but not suitable acres come from past timber harvest activities, with wildfire being the second largest source of capable but currently unsuitable habitat on the Umpqua NF.

36

Table 5. Environmental Baseline: Umpqua National Forest environmental baseline as of March 2013, utilizing the 15 year monitoring report for the Northwest Forest Plan spotted owl habitat model (Davis 2011). NRF CAPABLE PROTECTED 1 UMPQUA NATIONAL FOREST – HABITAT HABITAT ACRES UNPROTECTED NORTHWEST FOREST PLAN ACRES ACRES (% OF ACRES ACRES (% TOTAL) (% TOTAL) TOTAL) (% OF TOTAL) 1. OWNERSHIP -All Ownerships 535,281 892,526 1,036,490 (52%) (86%) -Private 15,540 45,372 50,348 (31%) (90%) -Federal 519,741 890,302 579,581 406,579 986,135 (53%) (90%) (59%) (41%) 2. LAND ALLOCATIONS- Federal (hierarchal, no acres double-counted) -Congressionally Reserved 38,894 64,176 108,772 Areas (Wilderness, OCRA) (36%) (59%) -Late-Successional Reserves 159,747 278,115 279,553 (mapped, without IRA’s) (57%) (99.5%) - 100-Acre Spotted Owl 16,070 19,492 19,590 Core Areas in the Matrix (82%) (99.5%) - Areas Withdrawn from the timber base (Wild and Scenic River Corridor, 80,225 110,102 Inventoried Roadless 128,026 (63%) (86%) Areas(IRA’s), Administratively withdrawn) -Riparian Reserves (Matrix 51,955 96,667 98,640 and AMA acres only) (53%) (98%) -Matrix/Adaptive 172,850 323,429 Management Areas (IRA’s 351,554 (49%) (92%) removed) 3. 2012 FINAL CRITICAL HABITAT -Critical Habitat 297,455 313,631 596,633 (50%) (50%) -Critical Habitat, NRF 186,825 172,454 359,280 Habitat (52%) (48%) -Critical Habitat, Dispersal 49,099 106,738 57,639 (54%) Only Habitat (46%) -Critical Habitat, Capable 270,977 264,782 535,738 Habitat (51%) (49%) Notes: 1. Protected = land allocation with no programmed timber harvest which includes Congressionally Reserved land, LSR’s, 100 acre Owl Cores, Wild and Scenic River Corridors and Inventoried Roadless Areas.

37

Figure 6. Map of the Northwest Forest Plan Land Use Allocations on and adjacent to the Umpqua National Forest and the Lower Steamboat Timber Sale and Lonely Oak Project, as well as the 2012 Critical Habitat Sub-Unit boundaries.

38

III. Description of the Projects Being Analyzed for this Biological Assessment The action area is defined as "all areas to be affected directly or indirectly by the federal action and not merely the immediate area involved in the action.” (50 CFR 402). For this consultation, because the NSO is the only listed or proposed species likely affected by the preferred alternative, the action area includes all federal, state, and privately owned lands within 1.2 miles of proposed treatment areas and all known or potential NSO home ranges that could be directly, indirectly or cumulatively impacted through the preferred alternative.

A. Lower Steamboat Timber Sale

i. Lower Steamboat Timber Sale Action Area The NSO analysis area for the Loafer Timber Sale action area covers about 42,200 acres, of which 25,320 acres (60%) are currently suitable nesting, roosting, and foraging habitat (NRF) for the NSO (Table 6). There are 16 potential NSO home ranges within the analysis area. Approximately 12,238 acres of the action area (29%) are considered to be dispersal habitat. Less than half of the action area (18,194 acres) falls within 2012 critical habitat, Western Cascades South Sub Unit 5 (USFWS 2012, Figure 1). The Lower Steamboat planning area is located in the Middle North Umpqua and Lower Steamboat 5th-field watersheds within the Western Cascades physiographic province, and all units are located within Late Successional Reserve RO-222. Land ownership in the action area is mostly the Umpqua NF (40,339 acres), followed by private land (1,135 acres) and then BLM (666 acres) (Table 6, Figure 7). The Lower Steamboat action area is located primarily within three forest vegetation zones, including the White Fir Zone, Western Hemlock Zone, and Douglas-fir Zone. At high elevations, the planning area also includes small inclusions of the Mountain Hemlock Zone and Pacific Silver Fir Zone.

39

Figure 7. Map of the Lower Steamboat Timber Sale action area and harvest units, along with 2012 Northern Spotted Owl Critical Habitat and owl cores.

40

Table 6. Description of the total Action Area environmental baseline for the Lower Steamboat Timber Sale by land use allocation and amount of suitable and capable habitat present utilizing the 15 year monitoring report for the Northwest Forest Plan spotted owl habitat model (Davis 2011).

2 UMPQUA NATIONAL FOREST, NSO NRF CAPABLE NSO PROTECTED DISPERSAL 1 UNPROTECTED LOWER STEAMBOAT TIMBER HABITAT HABITAT ACRES ACRES ACRES ACRES SALE – NORTHWEST FOREST ACRES ACRES (% OF (% OF (% OF TOTAL) PLAN (% TOTAL) (% TOTAL) TOTAL) TOTAL) 1. OWNERSHIP 25,320 40,934 39,230 12,238 -All Ownerships 42,200 2,970 (7%) (60%) (97%) (93%) (29%) -Private, State, and other non-BLM 1,135 Government 24,813 40,244 39,230 1,835 36,883 -Federal 41,065 (60%) (98%) (93%) (45%) (90%) 2. LAND ALLOCATIONS- Federal (hierarchal, no acres double-counted) -Congressionally Reserved Areas (Oregon Cascades 1,070 1,103 746 (68%) 0 0 Recreation Area, (97%) Administratively Withdrawn) 39,230 -Late-Successional 22,964 37,998 (93%) 38,022 0 0 Reserves (mapped) (60%) (99%) - 100-Acre Spotted Owl Core Areas in 105 67 (60%) 104 (99%) the Matrix -Riparian Reserves (Matrix and AMA 324 165 (51%) 311 (96%) acres only) -Matrix/Adaptive Management 785 404 (52%) 761 (97%) Areas 3. FINAL 2012 CRITICAL HABITAT 2 NSO NRF CAPABLE NSO DISPERSAL HABITAT HABITAT ACRES ACRES ACRES ACRES (% OF (% TOTAL) (% TOTAL) TOTAL) -Final Critical Habitat 12,372 17,648 16,920 18,194 2012 (68%) (97%) (93%) Notes: 1. Protected = land allocation with no programmed timber harvest which includes Congressionally Reserved land, LSR’s, Owl Cores and Wild and Scenic River Corridors. 2. Dispersal does not include NRF habitat.

41

ii. Status of the Species in the Lower Steamboat Action Area There are no current surveys for owls within the action area. Due to the lack of recent survey data, spotted owl activity centers and home ranges within the action area have been estimated using the method identified by USDI-USDA (2008). That analysis resulted in an estimate of 16 potential spotted owl home ranges/territories located around activity centers based either on known historic or on predicted activity center locations (KPACs) within the action area (Figure 8 and Table 7). Of these 16 KPACs, 11 are based upon historic owl pair locations identified from surveys which occurred in the late 1990’s, and the remaining 5 were located as per USDI-USDA (2008).

Figure 8. Lower Steamboat timber sale action area and all associated owl sites within the action area, along with the Nesting Roosting and Foraging (NRF) habitat layer.

For the analysis for this consultation, the analysis of spotted owl home range/territory condition was conducted using the 1.2 mile provincial home range radius (for the Oregon Cascades province) around each KPAC. The current condition of the spotted owl home ranges/territories associated with the KPACs within the action area is described in Table 7. All of the KPACs are above the 40% home range and 50% core use area thresholds for amounts for NRF habitat.

42

Table 7. Northern Spotted Owl activity centers within the Lower Steamboat Action Area listed by ID number, with percent of nesting, roosting and foraging (NRF) habitat at the 800m core scale and 1.2 mile home range scale.

ID SITE NAME STATUS Core % NRF Home Range % NRF

374 COMPUTED 71% 68% 387 COMPUTED 85% 78% 419 COMPUTED 76% 65% 432 COMPUTED 77% 57% 564 DOG CREEK PAIR 63% 67% 565 WINSLOW PAIR 70% 59% 568 JACK CREEK PAIR 72% 58% 577 UPPER SINGE CREEK PAIR 67% 66% 583 LITTLE FALLS RESIDENT SINGLE 66% 60% 585 DEEP CREEK PAIR 62% 51% 586 SINGE CREEK PAIR 74% 70% 587 BLACK GORGE PAIR 51% 45% 594 ILLAHEE PAIR 59% 60% 595 REYNOLDS RIDGE RESIDENT SINGLE 58% 52% 596 SCAREDMAN PAIR 59% 44% 599 STEAMBOAT FALLS RESIDENT SINGLE 63% 55%

iii. Lower Steamboat Timber Sale Proposed Action Description The Purpose and Need of the Lower Steamboat Timber Sale is to promote the development and maintenance of late-successional forest conditions in existing even-aged stands in LSR (USDA & USDI, 1994b, C-12). Proposed project activities include silvicultural treatments designed to develop structurally complex stand and landscape structure and species composition within second-growth stands that originated following even-aged management and subsequent planting in the 1950s through the 1970s. A reduction in some natural and activity-generated fuels is an additional beneficial outcome associated with the project.

Stands proposed for silvicultural treatments within the project area are relatively homogeneous, even- aged, Douglas-fir stands ranging from 38 to 57 years old. These stem exclusion stands were established following even-aged harvest in the 1950s through 1970s and subsequent planting. Relative to historic conditions, existing stands are characterized by limited species diversity and reduced structural complexity, with few legacy habitat components, such as large snags, large downed wood, and “wolf trees” (Spies & Franklin, 1991). In contrast to historic stem exclusion stand development patterns, plantations in the planning area are growing at much higher densities and experiencing pronounced intertree competition. While some of these stands could develop old-growth characteristics without silvicultural intervention, current stocking and structure of most of these stands were established to produce high timber yields, not to provide for the development of old-growth forests. Total acres of treatments and treatment types are shown in Table 8.

The proposed action treatments include: • Commercially thinning approximately 1,192 acres of dispersal habitat utilizing a range of silvicultural prescriptions that would generally retain approximately 40-100 trees per acre (tpa) 43

(Table 12). This range allows prescriptions to be applied to each unit depending on slope, aspect, and other site-specific conditions in order to increase growth, health, and vigor of the leave trees. Creation of ½-acre and 1-acre gaps also is proposed for 112 acres within treatment units; the prescriptions are designed to increase growth, health, and vigor of the leave trees and are anticipated to result in approximately 14.0 million board feet of timber. Activities described above include treatment within critical habitat, consisting of 508 acres of thinning, 236 acres of underburning thinned units and 66.5 acres of ½ to 1 acre gaps being created within current dispersal habitat (Table 11).

• Approximately 377 acres of ground-based, or mechanized, logging systems and 835 acres of skyline logging systems would be employed to implement the thinning prescriptions. Mechanized equipment is generally utilized on slopes under 35% and skyline systems over 35% slope.

• Approximately 200 landings are proposed, 160 of which are in skyline units and 40 in ground- based units. The average landing size is 0.05 acres in skyline units and 0.12 acres in ground- based units.

• Riparian reserves along fish-bearing (class 1 and 2) streams will not be thinned. Approximately 245 acres of riparian reserves along perennial non-fish bearing (class 3) and intermittent (class 4) streams are proposed for thinning treatments outside of the no-cut buffers. Class 3 no-cut buffers would be 85 feet each side of stream channel; class 4 no-cut buffers would be 25 feet each side of stream channel; where instability or slope breaks are present, buffers would be widened to protect these areas. Approximately 10 skyline landings are proposed within riparian reserves.

• Generally, felled material down to a six inch diameter top would be yarded and removed from the site and material from six to three inch diameter tops would be brought to the landings. Whole-tree yarding could occur, provided enough slash remains on site to meet temporary spur road obliteration and winterization requirements. Yarded material may be chipped, left on the landing for firewood cutters, or burned.

• The activity fuels, or slash, would be treated in order to break up the continuity of the fuels throughout the units. Grapple pile burning would occur on approximately 380 acres in the ground-based units. Areas with light fuel loading may be in a condition or location that would not benefit from applying a fuels treatment. Prescribed burning would occur on approximately 90 acres and 2 miles of handline would be constructed to support that burning. Landing piles would be burned. Implementation of these treatments would be subject to a post-harvest fuels assessment.

• Implementing a project-level amendment to the 1990 Umpqua National Forest Land and Resource Management Plan that would allow timber harvest within 150 feet of unique habitats. Timber harvest activities within these areas would not normally occur under the Standards and Guidelines for unique habitats (Wildlife Prescription C5-1, LRMP IV-200), but is considered beneficial for enhancing development of desired stand conditions.

• Implementing road work: no new system roads would be constructed and all temporary spur roads would be obliterated after use.

44

• Temporary spur road construction- approximately 0.4 miles of new temporary spur road would be constructed to gain access into thinning units, 50 feet of which would be located within riparian reserve areas, none would be located within no harvest buffers.

• Temporary spur road reconstruction- approximately 6 miles of temporary spur route to gain access into thinning units would be located on the existing footprint of skid roads, fire lines, and abandoned or unclassified roads that were built to access the original harvest units. Approximately 1.08 miles of these would be located within riparian reserve areas. Approximately 0.1 miles would be located in no-harvest buffers. Reconstruction will give the Forest Service the opportunity to properly obliterate and hydrologically restore these roads.

• Temporary spur road obliteration – after use, approximately 6.4 miles (combination of new and existing) of temporary roads would be obliterated with an excavator equipped with a “winged ripper” to de-compact soil as needed. Any excavated material, including soil and woody material, would be pulled back over the road. A native forage seed mix will be applied to all subsoiled temporary roads and landings to minimize erosion and the establishment of invasive weeds.

• System Road Reconstruction- Road reconstruction would include reconstruction to meet standards and guidelines of the Northwest Forest Plan, in order to accommodate flood flows, minimize the disruption of natural water flow pathways, and lessen the risk of erosion (ROD C 32-33), while providing for safe, cost-efficient timber haul.

Connected Actions • Steamboat Helibase: Manage vegetation to restore access to helipad #3 to facilitate emergency fire management and prescribed burning in the LSR. This includes harvesting all merchantable vegetation over 20” DBH, cutting and leaving all trees over 20” DBH and cutting and scattering non-merchantable vegetation over approximately 3 acres.

• Bloody Point Rock Pit and Disposal Site: Manage vegetation to provide better access and organization for rock source material and the administrative disposal area. Invasive weeds are propagating at the disposal area and the current site configuration makes it difficult to separate the disposal site from the source material, increasing the risk of contaminating the source material and spreading invasive weeds. The rock pit would also be expanded approximately 100’ to the north. This includes harvesting all merchantable vegetation over 20” DBH, cutting and leaving all trees over 20” DBH and cutting and scattering non-merchantable vegetation over approximately 2.7 acres.

•Similar and connected actions would include noxious weed treatments, watershed improvement projects, trail maintenance and repair activities, and fisheries improvement projects.

Table 8. Acres of proposed treatment by fuels and silvicultural prescription with 2012 Northern Spotted Owl Critical Habitat. Residual Trees Per Acre Piling Slash Underburning 40-70 TPA 238 236

45

70-100 TPA 34 0 Total Acres of Treatment In CHU 508 Total Acres of Gaps 66.5

Table 9. Silviculture and fuels treatment acres, pre and post treatment canopy covers, and amounts and sizes of gaps proposed as part of the Lower Steamboat Timber Sale. Silviculture and Fuels Total Acres Total Treatments in Trees Per Total Acres of Pre-Treatment Post-Treatment 1/2 Acre Acres 1 Acre (TPA) Treatment Canopy Cover Canopy Cover Gaps Acre Gaps 20 TPA (Bloody Point Rock Pit) 4 48% 19% 0 0 40-70 TPA, various fuels treatments 745 61% 40% 44 19 40-70 TPA, underburning 350 61% 35% 20 11 70-100 TPA 93 61% 46% 7 2 Total 1,192 71 32

B. Lonely Oak Project

i. Lonely Oak Project Action Area The NSO action for the Lonely Oak Timber Sale action area covers about 9,255 acres, of which 5,553 acres (60%) are currently suitable nesting, roosting, and foraging habitat (NRF) (Table 10, Figure 9). There are three potential NSO home ranges within the analysis area. Approximately 2,406 acres of the action area (26%) is considered to be dispersal habitat. The majority of the action area (5,228 acres) including all treatment units falls within the proposed 2012 critical habitat, Western Cascades South Sub Unit 5 (USFWS 2012, Figure 9). The project is entirely within the Clearwater 5th Field Watershed, in Matrix land use allocation as identified within the Northwest Forest Plan (USDA/USDI 1994). The project is located on the Diamond Lake Ranger District.

The action area consists of a mixture of mature dominant and co-dominant Douglas-fir, ponderosa pine, sugar pine, and incense cedar forested stands. They are uneven ages with small groups of Douglas and white fir regeneration. Much of the overstory Douglas-fir has conk present. Ocean spray, salal, and dwarf Oregon-grape are common in the understory. The southern portion of the treatment units includes a rocky outcrop which contains a few oaks clinging to the top of the cliffs overlooking the Clearwater River. Directly below the treatment area is the canal system for the North Umpqua PacifiCorp hydropower system.

46

Figure 9. Lonely Oak Project action area and treatment units, along with 2012 Northern Spotted Owl Critical Habitat and 100 acre late successional reserve owl cores.

47

Table 10. Description of the total Action Area environmental baseline for the Lonely Oak Timber Sale by land use allocation and amount of suitable and capable habitat present utilizing the 15 year monitoring report for the Northwest Forest Plan spotted owl habitat model (Davis 2011).

NSO CAPABLE 2 NRF NSO PROTECTED DISPERSAL 1 UNPROTECTED HABITAT HABITAT ACRES ACRES ACRES ACRES UMPQUA NATIONAL FOREST, ACRES ACRES (% OF (% OF (% OF TOTAL) LONELY OAK TIMBER SALE – (% (% TOTAL) TOTAL) NORTHWEST FOREST PLAN TOTAL) TOTAL) 1. OWNERSHIP 4,362 7,437 2,125 -All Ownerships 7,589 267 (4%) 7,322 (96%) (57%) (98%) (28%) -Private, State, and other 0 0 0 0 0 0 non-BLM Government 7,589 4,362 7,437 267 (4%) 7,322 (96%) 2,125 -Federal (57%) (98%) (28%)

2. LAND ALLOCATIONS- Federal (hierarchal, no acres double-counted) -Congressionally Reserved Areas (Oregon Cascades Recreation Area, 0 0 0 0 0 Administratively Withdrawn) -Late-Successional 0 0 0 0 0 Reserves (mapped) - 100-Acre Spotted Owl 231 267 267 0 36 (13%) Core Areas in the Matrix (87%) (100%) -Riparian Reserves 713 1,064 (Matrix and AMA acres 1,097 0 0 263 (24%) (65%) (97%) only) -Matrix/Adaptive 3,382 6,101 1,826 6,225 0 0 Management Areas (54%) (98%) (29%) 3. FINAL 2012 CRITICAL HABITAT NSO CAPABLE 2 NRF NSO DISPERSAL HABITAT HABITAT ACRES ACRES ACRES ACRES (% OF (% (% TOTAL) TOTAL) TOTAL) -Final Critical Habitat 2,387 3,491 1,104 3,562 221 (6%) 3,341 (94%) 2012 (67%) (98%) (31%) Notes: 1. Protected = land allocation with no programmed timber harvest which includes Congressionally Reserved land, LSR’s, Owl Cores and Wild and Scenic River Corridors. 2. Dispersal includes NRF habitat.

48

ii. Status of the Species in the Lonely Oak Action Area There are no current surveys for owls within the action area. Due to the lack of recent survey data, spotted owl activity centers and home ranges within the action area have been estimated using the method identified by USDI-USDA (2008). That analysis resulted in an estimate of 3 potential spotted owl home ranges/territories located around activity centers based either on known historic or on predicted activity center locations (KPACs) within the action area (Figure 10 and Table 11). Of these 3 KPACs, one is based upon a historic owl pair location identified from surveys which occurred in the late 1990’s, and the remaining 2 were located as per USDI-USDA (2008). For the analysis for this consultation, the analysis of spotted owl home range/territory condition was conducted using the 1.2 mile provincial home range radius (for the Oregon Cascades province) around each KPAC. The current condition of the spotted owl home ranges/territories associated with the KPACs within the action area is described in Table 11. All of the KPACs are above the 40% home range and 50% core use area thresholds for amounts for NRF habitat.

Figure 10. Lonely Oak project action area and all associated owl sites within the action area, along with the Nesting Roosting and Foraging (NRF) habitat layer.

Table 11. Northern Spotted Owl activity centers within the Lower Steamboat Action Area listed by ID number, with percent of nesting, roosting and foraging (NRF) habitat at the 800m core scale and 1.2 mile home range scale. Core % Home Range ID SITE NAME STATUS NRF % NRF 349 Computed 72% 56% 397 Computed 72% 65% 539 Mowich Pair 87% 58%

49

iii. Lonely Oak Project Proposed Action Description The Lonely Oak Project has four primary purpose and need elements: 1. Creation or restoration of unique wildlife habitats (Oregon white oak, dry forest/shrublands, and wetlands); 2. Releasing ponderosa and sugar pines to increase their vigor and resiliency by removing encroaching conifers; 3. Reintroducing fire as a natural disturbance element and reducing fuel loads; 4. Creation of merchantable wood-related products to provide jobs and income to the local economy. Project purposes are the result of land and resource management needs identified in the State of Oregon’s Conservation Strategy, the Lower Clearwater Watershed Assessment and the Umpqua Forest’s 2003 Watershed Restoration Business Plan. The Oregon Department of Fish and Wildlife’s Oregon Conservation Strategy identifies oak woodlands as one of the priority habitats for restoration in the West Cascades ecoregion. This analysis recommends the use of multiple tools, including prescribed fire and selective harvest to manage oak dominated woodlands. The Lonely Oak project consists of a total of 127 acres of treatments. The treatments are a combination of prescribed fire and silvicultural treatments. Stands along the southern boundary of the project area will be thinned to approximately 40% canopy cover in order to reduce conifer encroachment around legacy ponderosa pine and to open up stand conditions to be able to plant Oregon white oak post treatment. The remaining stands will be thinned to release ponderosa pine and sugar pines and allow for more sunlight penetration to the forest floor to develop a more robust herbaceous understory for ungulate forage. All treatment units will be underburned post-harvest, and there are approximately 51 acres of prescribed fire only treatment (Table 12). Table 12. Acres of treatment proposed as part of the Lonely Oak Project.

Treatment Total Acres of Pre-Treatment Post-Treatment Treatment Canopy Cover Canopy Cover Prescribed Fire Only 51 91% 69% Wetland Creation 2 85% 85% Thin to ~61% canopy cover for pine release 3.5 77% 61% Thin to ~40% canopy cover, plant with oak 70.5 77% 41% Total Acres 127

Activities proposed in the proposed CHU, include 35 acres of underburning and non-commercial thinning, 60 acres of commercial thinning, and 2 acres of wetland creation (with no overstory removal) (Table 13, Figure 10).

Table 13. Acres by treatment type within 2012 Northern Spotted Owl Critical Habitat that is part of the Lonely Oak Project. Treatment Type Acres Non Commercial Thin and underburn 27.6 Oak/Pine retention, underburn 0 Thin to 120-160 tpa 57 Underburn Only 6.8 Wetland Creation, excavation and spoil 2

50

site Total Acres 93.4

Connected Actions Native Species Revegetation Treatments All of Units 2 and 5, and portions of Units 1, 4 and 6 may be revegetated with native species. In Units 1 and 2, broadleaved trees (Oregon white oak Quercus garryana, madrone Arbutus menziesii, and chinquapin Chrysolepis chrysophylla) will be planted to augment the availability of this unique habitat type. Unit 4 includes the location of a small (less than 1/2 acre) wetland. This site would also be revegetated with wetland appropriate native species. All of unit 5 may also be seeded with herbaceous, native species that have desired big game forage qualities. Unit 6 has a prescribed burn prescription. In the event there are areas of high intensity burn greater than 1/10th of an acre, these areas would be seeded to re-establish a vegetative ground cover. Portions of Units 2 & 5 may also be treated with fungal inoculant to enhance desired fungal species for wildlife and human use.

Maintenance Lopping and Underburning To maintain open site conditions in Units 2 and 5, mechanical removal of conifer seedlings may be required. Maintenance of stand conditions throughout the entire project footprint may also be accomplished through prescribed burning at periodic intervals.

Noxious Weed Control Treatments As forest canopy is reduced, the treatment locations have an increased susceptibility to invasion by noxious weeds. Post project site monitoring would be conducted to assess treatment needs. Likely control methods include hand-pulling and solarization. All units of the alternative may be considered for this treatment, but anticipated treatment levels are quite low.

IV. Direct And Indirect Effects

The direct and indirect effects are analyzed at two scales, 1) stand-scale and 2) the action area.

A. Lower Steamboat

Effects to spotted owl NRF habitat The Lower Steamboat Timber Sale units nor any of the new temporary roads being proposed occur within NRF habitat, therefore there will be no direct effects to NRF habitat. Indirect effects to NRF habitat may occur through the use of skyline logging systems which may require tailhold trees located within NRF habitat. All tailhold trees outside of timber sale units within NRF habitat will be left on site, and left standing if safe to do so. Effects to spotted owl Dispersal habitat All of the 1,192 acres of treatment proposed as part of the Lower Steamboat Timber Sale are currently dispersal habitat. Post treatment, 802 acres of dispersal habitat will be retained above 40% canopy cover, while 392 acres of dispersal habitat will be removed as a result of gap creation and combination silviculture prescriptions and underburning which will reduce canopy covers on 682 acres to 35% canopy cover, with an additional 63 acres of removal as a result of within unit gaps. Additionally, with the creation of 4/10ths of a mile of new temporary road, an additional .7 acres of habitat will be downgraded. The other temporary roads will be built upon previous temporary roads which are currently not dispersal habitat due to the young, <30ft trees growing on the existing disturbance. Finally, 51 as part of the expansion of the Bloody Point Rock Pit, an additional 4 acres of dispersal habitat will be removed. The totals amounts of dispersal habitat impacted are listed in Table 14. Table 14. Table of impacts to dispersal habitat resulting from the Lower Steamboat Timber Sale. Dispersal Habitat Removed from Dispersal Habitat Dispersal Habitat Treatment Gaps within Units of Dispersal Removed from Entire Maintained Habitat Maintained Units Thinnings 802 40 352 Temp Road 0 0 .4 Construction Totals 802 40 352.4

Impacts to individual owl home ranges, core use areas and nest patches are described in Table 15. Five KPACs have reductions in dispersal habitat at the home range and core use area scales. The largest reduction is in the Jack Creek site where the amount of dispersal habitat is reduced by 7% at the home range scale and 10% at the core use area scale. Table 15. Table of all owl sites amount of dispersal habitat pre and post treatment within the Lower Steamboat Timber Sale action area. Pre Treatment Dispersal Post Treatment Dispersal Habitat Habitat Home Nest Home Nest ID SITE NAME STATUS Range Core Patch Range Core Patch 374 COMPUTED 27% 27% 39% 27% 27% 39% 387 COMPUTED 14% 13% 19% 14% 13% 19% 419 COMPUTED 27% 21% 18% 27% 21% 18% 432 COMPUTED 31% 18% 5% 31% 18% 5% 564 DOG CREEK PAIR 25% 27% 8% 24% 22% 7% 565 WINSLOW PAIR 32% 21% 2% 32% 21% 2% 568 JACK CREEK PAIR 33% 24% 3% 26% 14% 3% 577 UPPER SINGE CREEK PAIR 27% 32% 9% 25% 30% 9% RESIDENT 583 LITTLE FALLS SINGLE 32% 29% 20% 29% 28% 20% 585 DEEP CREEK PAIR 34% 28% 7% 33% 25% 7% 586 SINGE CREEK PAIR 21% 19% 7% 21% 19% 7% 587 BLACK GORGE PAIR 36% 29% 17% 36% 29% 17% 594 ILLAHEE PAIR 29% 27% 18% 27% 27% 18% RESIDENT 595 REYNOLDS RIDGE SINGLE 38% 21% 34% 38% 21% 34% 596 SCAREDMAN PAIR 44% 40% 16% 44% 40% 16% RESIDENT 599 STEAMBOAT FALLS SINGLE 36% 29% 4% 36% 29% 4%

Effects to spotted owl prey habitat The effects to spotted owl prey will be combined for both projects and will be discussed later.

52

Effects to Spotted Owl Stand and Landscape level Dispersal The Lower Steamboat action area has a total of 25,320 acres of suitable NRF habitat, which would be maintained as a result of the implementation of the proposed project. There are 1,192 acres of treatment scattered across the 42,200 acre action area, so less than 3% of the action area will be treated. The action area has 12,238 acres of dispersal habitat, of which 392 acres will be downgraded from dispersal, which is 3% of the total dispersal habitat in the action area meaning that 97% of the dispersal habitat will be maintained after the implementation of this project.

The effects of the proposed treatments across the landscape will be beneficial to spotted owls in the long term as the stands being treated will develop more structural diversity horizontally and vertically as a result of the variable density thinning and scattered ½ to 1 acre gaps. As the treated stands are all within LSR, under the NWFP they are not likely to be harvested in the future and will be allowed to develop into suitable NRF habitat over a time period of 30-50 years (Wessell-Kelly Forest Vegetation Analysis for Lower Steamboat). There may also be additional landscape level benefits in that potential future wildfires that burn through more heterogeneous stands often burn at lower intensities than they do in single storied, tight canopied younger forest (Umpqua NF 2003).

Anchor and Hazard Tree Removal There will be site-specific adverse effects to NRF habitat where anchor (tailhold and guyline) trees and hazard trees associated with harvest units (anchor/hazard trees) are felled as part of the Lower Steamboat project. There are harvest units (2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 14, 16, 18, 19, 20, 21, 23, 24, 100, 103, 104, 106, 108, 111, 112, 113, 115, 116, 211 and 212) that are adjacent to mapped spotted owl NRF habitat that will utilize skyline logging and may require the use of anchor trees. Guyline trees will be located within the units whenever possible. Whereas the guyline trees within units will be felled, tailhold trees will be left standing when feasible and the smallest possible trees will be selected, but trees of all diameters could be used. Use of old-growth trees will be avoided if possible. No guyline trees are proposed for felling within 100-acre spotted owl LSRs. Felled guyline trees will not continue to function as potential nest or roost trees in these stands, but they will remain on site to serve as habitat for prey species. Anchor Tree Removal will avoid, when possible: • Known spotted owl nest trees, or adjacent trees providing habitat function; • Trees with nesting structure for spotted owls, or adjacent trees providing habitat function. Although the availability of potential nest trees might not appear to be the factor limiting spotted owl use of the affected stands, site-specific spotted owl surveys are not part of the proposed action.

Therefore, a reasonable level of uncertainty remains regarding the role(s) identified anchor/hazard trees might be playing in spotted owl habitat quality within the affected stands. For example, without surveys, we are unable to determine with certainty where any spotted owls using the stand might be nesting; the potential for removing nest trees remains. Nor, will one always be able to determine from the ground whether the identified anchor/hazard tree in question harbors a communal den. Because nest sites could still be removed, because loss of communal den sites seems likely to reduce the flying squirrel carrying capacity at the stand scale, and further because some studies have found evidence of positive correlations between large prey (flying squirrel and woodrat) availability and spotted owl reproductive success (Barrows 1987, Laymon 1988, Thrailkill and Bias 1989, but see White 1996, Forsman et al. 2001, 2004), the removal of anchor/hazard trees will, in some cases, affect stand-level habitat quality/suitability. Although trees with nesting structure for spotted owls will be avoided when possible, it will not always be possible to avoid them – thus, nest trees or adjacent trees providing habitat function could be removed. Therefore, under certain circumstances, anchor/hazard tree removal will likely adversely affect spotted owls.

53

Effects to Spotted Owls due to Disturbance The proposed project may also have direct effects on NSOs through noise generating disturbances within close proximity to both known NSO activity centers and spatially suitable habitat that may support nesting owls. It is expected that some potential disturbance activities would occur during the March 1 through September 30 NSO breeding season. Activities occurring after the critical breeding period (March 1 through July 15) may disturb the NSO, but are not likely to disrupt NSO reproductive success because the identified buffers and restrictions from the programmatic disturbance letter of concurrence (USFWS 2009, TAILS 13420-2009-I-0070) are being applied (Table 16, Figure 11). There will be full seasonal restrictions placed upon the connected action of the expansion of the Bloody Point Rock Pit as this could involve explosives (March 1 through September 30). For this project land managers would like to have the flexibility of burning in the spring or in the summer to increase the likelihood of hitting a suitable burn window. Past experience on the Umpqua in implementing fall burning has demonstrated that hitting suitable fuel moisture and climate conditions suitable for burning in the fall alone are very difficult to meet. Therefore the option of burning within the spring is desireable. Smoke and fire personnel activity are likely to adversely affect any spotted owls utilizing the area during burning operations in the spring. Due to the low intensity prescriptions and the desire for low intensity fire there will not be any anticipated crown fire, and desired effects will be low flame lengths traveling along the forest floor.

Table 16. Disruption distances for the northern spotted owl during the breeding period of March 1- September 30.

Disruption Distance

Disturbance Source Critical Breeding Period Latter Breeding Period (July 16 – (March 1 – July 15) September 30)

Aircraft – Type I Helicopters 402 meters (0.25 mile) 402 meters (0.25 mile) Aircraft – Fixed wing and other3 110 meters (120 yards) None Helicopters Blasting – Quarries; more than two 1610 meters (1 mile) 402 meters (0.25 mile) pounds of explosives Blasting – Small charge; two pounds 110 meters (120 yards) None or less of explosives Campground Maintenance - Heavy 60 meters None Chainsaw use - General 60 meters (65 yards) None Heavy equipment 32 meters (35 yards) None Invasive Plant Control – Power 60 meters None Equipment Level 2 Road and Dike Repair: Heavy 60 meters None equipment, pile driving

54

Disruption Distance

Disturbance Source Critical Breeding Period Latter Breeding Period (July 16 – (March 1 – July 15) September 30) Precommercial Thinning and Pruning 60 meters None Prescribed Burning 402 meters (0.25 mile) None Quarry Operations: Rock crushing, 165 meters (180 yards) None drilling, hauling, etc. Road Maintenance/Brushing 32 meters (35 yards) None

Timber harvesting and road construction associated activities (e.g. temporary road construction, culvert replacement, chainsaw and etc.) produce above ambient noise levels that have the potential to disturb nesting NSOs, and disrupt normal reproductive activities. Disturbance could cause nest abandonment, flushing of adults off eggs thereby exposing them to harm, depressed feeding rates and avoidance of otherwise suitable habitat.

Figure 11. Lower Steamboat Timber Sale disturbance map showing the 60m and quarter mile disturbance buffers around all KPAC’s within the action area.

55

B. Lonely Oak

Effects to spotted owl NRF habitat The Lonely Oak Project will result in the downgrading of 69 acres of currently suitable NRF habitat to dispersal habitat, of which 45 acres are within critical habitat (Table 17). The 69 acres of downgraded NRF habitat are less than 2% of the total NRF habitat within the action area. The treatments will retain the largest trees, and the treatments are in part designed to release encroaching conifers from around ponderosa and sugar pines. All treatment units will also be treated with prescribed fire which will likely cause limited mortality of overstory and midstory trees. Prescribed fire will also likely consume some snags and create additional snags within the treated areas. Underburning has been targeted specifically for these units as a restorative treatment due to the higher density of fire dependent ponderosa pines and lower densities of sugar pines as well as the south aspects of these units. The intent of the underburning is twofold: primarily it is designed to reduce surface and ladder fuels, while at the same time reintroducing fire to stands that based upon the predominance of ponderosa pine within them would have been more likely to have more frequent lower intensity fire within them which the proposed underburning is attempting to mimic.

Table 17. Amount of NRF and Dispersal habitat maintained and amount of NRF habitat downgraded to dispersal habitat as a result of the Lonely Oak project. Dispersal Habitat Maintained NRF Habitat Maintained NRF Habitat Downgraded to Dispersal 58 6 69

All of the existing KPAC’s within the action area are currently above threshold levels for amounts of NRF habitat at both the home range and core use area scale (Table 18). Post treatment, the KPAC’s will remain above thresholds for NRF habitat at both the home range and core use area scale. However there will be a reduction of 2.5% NRF habitat (69 acres) at the home range scale and a reduction of 14% (69 acres) for KPAC # 397. There will be 7 acres of underburning within the nest patch of this KPAC, but it will not remove overstory canopy, and therefore the habitat will be maintained. Table 18. Amounts of nesting, roosting and foraging habitat at the home range (1.2 mile), core (800m) and nest patch (70m) scale pre and post Lonely Oak Project treatments for the KPAC’s within the Lonely Oak Project Action area. Pre-treatment NRF Post treatment NRF SITE Home Nest Nest ID STATUS Core Home Range Core NAME Range Patch Patch 349 Computed 56% 72% 84% 56% 72% 84%

62.5% (-2.5% or 58% (-14% or 397 Computed 65% 72% 85% 85% 69 acres) 69 acres) 539 Mowich Pair 58% 87% 96% 58% 87% 96%

Effects to spotted owl Dispersal habitat There are 127 total acres of treatment proposed as part of the Lonely Oak Project, of which 58 acres are dispersal habitat. Post treatment all 58 acres of dispersal habitat will be retained above 40% canopy cover, and therefore will be maintained as dispersal habitat. There are no new temp roads being proposed as part of the Lonely Oak project. The proposed for treatment that will maintain dispersal

56 habitat may be planted with native hardwood species like Oregon white oak, madrone or chinquapin. Long-term, this planting will develop into an additional canopy story and will produce food for spotted owl prey items.

The creation of the small wetland will occur within current dispersal habitat and no overstory will be removed as a result of the construction of the wetland. The wetland will be ½ acre in size and will remove vine maple within the wetland footprint. It will not be plumbed; it will be filled with overwinter rain and snow. This will not change the overall canopy cover of the stand as no overstory trees will be removed.

Dispersal habitat will also be underburned, and will experience the same impacts as described in the above section on underburning within NRF habitat.

Anchor and Hazard Tree Removal There is no skyline harvest proposed as part of the Lonely Oak project, it will all be ground based treatment. Therefore there will be no anchor tree or guyline tree use as part of this project.

Effects to Spotted Owl Stand and Landscape level Dispersal The Lonely Oak action area has a total of 4,362 acres of suitable NRF habitat, of which 4,293 acres would be maintained as a result of the implementation of the proposed project, resulting in the reduction of less than 2% of the NRF habitat within the action area. The action area has 2,125 acres of dispersal habitat, all of which will be maintained after the implementation of this project. Due to the small amount of the action area being treated (less than 2%) there will be limited impacts to spotted owl dispersal as a result of the Lonely Oak Project.

Effects to Spotted Owls due to Disturbance The Lonely Oak project may also have direct effects on NSOs through noise generating disturbances within close proximity to both known NSO activity centers and spatially suitable habitat that may support nesting owls. It is expected that some potential disturbance activities would occur during the March 1 through September 30 NSO breeding season. Activities occurring after the critical breeding period (March 1 through July 15) may disturb the NSO, but are not likely to disrupt NSO reproductive success because the identified buffers and restrictions from the programmatic disturbance letter of concurrence (USFWS 2009, TAILS 13420-2009-I-0070) are being applied (Table 16, Figure 12). For this project land managers would like to have the flexibility of burning in the spring or in the summer to increase the likelihood of hitting a suitable burn window. Past experience on the Umpqua in implementing fall burning has demonstrated that hitting suitable fuel moisture and climate conditions suitable for burning in the fall alone are very difficult to meet. Therefore the option of burning within the spring is desirable. Smoke and fire personnel activity are likely to adversely affect any spotted owls utilizing the area during burning operations in the spring. Due to the low intensity prescriptions and the desire for low intensity fire there will not be any anticipated crown fire, and desired effects will be low flame lengths traveling along the forest floor.

57

Figure 12. Lonely Oak project disturbance map showing the 60m and quarter mile disturbance buffers around all KPAC’s within the action area.

Effects to spotted owl prey habitat for both projects Initial treatments to both dispersal and NRF habitat will affect prey habitat by reducing understory diversity, but may beneficially increase other prey items like Peromyscids in the short term (approximately 2 years post treatment) (Converse et al. 2006). Truffle populations are likely to decline post treatment as a result of the more open stand structure and decreased downed wood, which can have negative short term implications to foraging opportunities at the stand level (Meyer et al. 2005, Lehmkuhl et al. 2006). However, these short term impacts may be partially offset if the treated stands are burned in a wildfire, as the treated stands will likely retain more habitat post fire than untreated stands would post fire (Lehmkuhl et al. 2006).

Recent research on the effects of thinning on northern flying squirrel populations indicate that thinning which removes midstory canopy layers results in decreased flying squirrel populations (Wilson 2010, Manning et al. 2012, Wilson personal communication 2011), due, perhaps in part, to increased predation risk. These effects have been measured to last a minimum of 11 years post thinning, but may last upwards of 50 to 60 years depending upon the level of canopy removed as a result of the thinning and or the productivity of the site affecting the growth rate of trees (Wilson 2010, Manning et al. 2012, Wilson personal communication 2011). Manning et al. (2012) found a reduction of .75 flying squirrel/acre in the first year post light thinning compared to the control stands, and a reduction of .5 flying squirrel/acre in the second year post light thinning. Wilson (2010) found that flying squirrel populations in treated stands remained below control stands for 12 years after the initial treatment – the duration of their study. The initial reduction of flying squirrels is believed to be due to predation, as increased canopy spacing and loss of midstory canopy both increases gliding distances and required travel along the ground which results in high post treatment predation by both avian and mammalian

58 predators (Wilson 2010). This depression in flying squirrel populations in treated stands is expected to persist until suitably complex midstory canopy develops to support flying squirrel dispersal through the stand (Wilson pers. comm 2011).

Flying squirrels are the most prominent prey for spotted owls, but they will also prey upon a variety of other small mammals, particularly woodrat species (Forsman et al. 1984, 2004; Ward et al. 1998). While there may be woodrats are present within the action area, their densities are unknown. Underburning may adversely affect woodrat nests, but underburning generally results in a flush of herbaceous growth in the year following the burning which has a beneficial impact on deer mice populations which are a secondary spotted owl prey item.

Red tree voles (Arborimus longicaudus) are an additional prey item for spotted owls. They prefer older, more structurally diverse forests for nesting habitat, but they have been demonstrated to use younger forests with tight canopies for foraging and nesting (Wilson and Forsman 2013). Thinning these younger stands also increases tree vole susceptibility to predation as a result of the reduced canopy densities (Wilson and Forsman 2013). This happens because tree voles in younger stands depend upon multiple trees for food resources (conifer needles), so individuals must forage in multiple trees adjacent to their nests. The increased spacing of trees post thinning requires the tree voles to climb down the trees and run across the forest floor to the next tree, increasing their exposure to predators and increasing their energetic needs (Wilson and Forsman 2013).

Some of the rationale for gap creation is to create and accelerate mid-story canopy development. Long term, the proposed gap treatments and variable density thinning prescriptions proposed as part of the Lower Steamboat Timber Sale will increase structural diversity and development of older forest characteristics which will be beneficial to flying squirrels and tree voles (Wilson and Forsman 2013). Additionally, within and adjacent to treatment units in both projects there will areas of unthinned younger forest habitat that will not be treated and will provide refugia habitat, along with the adjacent older forest which will be beneficial for maintaining prey populations across the landscape. Due to adverse effects to prey habitat by increasing canopy spacing and potential short term loss of midstory canopies both the Lower Steamboat and Lonely Oak projects will adversely affect the northern spotted owl.

V. Direct and Indirect Effects to Critical Habitat

In December of 2012, the USFWS released a final rule designating critical habitat for the northern spotted owl (USFWS 2012a). The action area for Lower Steamboat contains 18,194 acres of revised critical habitat, all of which falls within the Western Cascades South sub unit 5 (Figure 10, Table 6). There are 508 acres of Lower Steamboat units within WCS-5. The action area for Lonely Oak contains 3,562 acres of critical habitat which also all falls within WCS-5, and 93.4 acres of treatment within WCS-5 (Figure 8, Table 12). The final rule states, “In general, prescriptions (e.g., vegetation management, prescribed fire, etc.) that apply ecological forestry principles to address the restoration and conservation of broader ecological processes in areas where this is needed, while minimizing impacts to structurally diverse or mature and old forest that does not require such management can be compatible with maintaining the critical habitat’s essential features in the long term at the landscape scale” (USFWS 2012a, p. 71882). The final rule also states that, “This rule is different from previous designations of northern spotted owl critical habitat in that we are recommending a ‘‘hands on’’ approach to forest management within critical habitat. We encourage land managers to consider active management of

59 forests that balance short-term impacts with long-term beneficial effects, which ultimately supports long-term conservation of the northern spotted owl” (USFWS 2012a, p. 72014).

The final rule on the revised 2012 critical habitat defined primary constituent elements (PCEs). The PCEs include: (1) Forest types that may be in early-, mid-, or late-seral stages and that support the northern spotted owl across its geographical range; these forest types are primarily: (a) Sitka spruce, (b) Western hemlock, (c) Mixed conifer and mixed evergreen, (d) Grand fir, (e) Pacific silver fir, (f) Douglas-fir, (g) White fir, (h) Shasta red fir, (i) Redwood/Douglas-fir (in coastal California and southwestern Oregon), and (j) The moist end of the ponderosa pine coniferous forests zones at elevations up to approximately 3,000 ft. (900 m) near the northern edge of the range and up to approximately 6,000 ft. (1,800 m) at the southern edge. (2) Habitat that provides for nesting and roosting. In many cases the same habitat also provide for foraging (PCE (3)). Nesting and roosting habitat provides structural features for nesting, protection from adverse weather conditions, and cover to reduce predation risks for adults and young. This PCE is found throughout the geographical range of the northern spotted owl, because stand structures at nest sites tend to vary little across the spotted owl‘s range. These habitats must provide: (a) Sufficient foraging habitat to meet the home range needs of territorial pairs of northern spotted owls throughout the year. (b) Stands for nesting and roosting that are generally characterized by: (i) Moderate to high canopy closure (60 to over 80 percent), (ii) Multilayered, multispecies canopies with large (20- 30 in (51-76 cm) or greater dbh) overstory trees, (iii) High basal area (greater than 240 ft2/acre (55 m2/ha)), (iv) High diversity of different diameters of trees, (v) High incidence of large live trees with various deformities (e.g., large cavities, broken tops, mistletoe infections, and other evidence of decadence), (vi) Large snags and large accumulations of fallen trees and other woody debris on the ground, and (vii) Sufficient open space below the canopy for northern spotted owls to fly. (3) Habitat that provides for foraging, which varies widely across the northern spotted owl‘s range, in accordance with ecological conditions and disturbance regimes that influence vegetation structure and prey species distributions. Across most of the owl‘s range, nesting and roosting habitat is also foraging habitat, but in some regions northern spotted owls may additionally use other habitat types for foraging as well. The specific foraging habitat PCEs for the four ecological zones within the geographical range of the northern spotted owl are the following: (a) West Cascades/Coast Ranges of Oregon and Washington (i) Stands of nesting and roosting habitat; additionally, owls may use

60

younger forests with some structural characteristics (legacy features) of old forests, hardwood forest patches, and edges between old forest and hardwoods; (ii) Moderate to high canopy closure (60 to over 80 percent); (iii) A diversity of tree diameters and heights; (iv) Increasing density of trees greater than or equal to 31 in (80 cm) dbh increases foraging habitat quality (especially above 12 trees per ac (30 trees per ha)); (v) Increasing density of trees 20 to 31 in (51 to 80 cm) dbh increases foraging habitat quality (especially above 24 trees per ac (60 trees per ha)); (vi) Increasing snag basal area, snag volume (the product of snag diameter, height, estimated top diameter, and including a taper function (North et al. 1999, p. 523)), and density of snags greater than 20 in (50 cm) dbh all contribute to increasing foraging habitat quality, especially above 4 snags per ac (10 snags per ha); (vii) Large accumulations of fallen trees and other woody debris on the ground; and (viii) Sufficient open space below the canopy for northern spotted owls to fly. (p. 71906-71907) (4) Habitat to support the transience and colonization phases of dispersal, which in all cases would optimally be composed of nesting, roosting, or foraging habitat (PCEs (2) or (3)), but which may also be composed of other forest types that occur between larger blocks of nesting, roosting, and foraging habitat. In cases where nesting, roosting, or foraging habitats are insufficient to provide for dispersing or nonbreeding owls, the specific dispersal habitat PCEs for the northern spotted owl may be provided by the following: (a) Habitat supporting the transience phase of dispersal, which includes: (i) Stands with adequate tree size and canopy cover to provide protection from avian predators and minimal foraging opportunities; in general this may include, but is not limited to, trees with at least 11 in (28 cm) dbh and a minimum 40 percent canopy cover; and (ii) Younger and less diverse forest stands than foraging habitat, such as even- aged, pole-sized stands, if such stands contain some roosting structures and foraging habitat to allow for temporary resting and feeding during the transience phase. (b) Habitat supporting the colonization phase of dispersal, which is generally equivalent to nesting, roosting, and foraging habitat as described in PCEs (2) and (3), but may be smaller in area than that needed to support nesting pairs.

The proposed thinning of dispersal habitat proposed as part of the Lower Steamboat Timber Sale will impact PCE (4 (a)) habitat that supports dispersal. However, only 66.5 acres of dispersal habitat will be downgraded in critical habitat, and no NRF habitat will be impacted. Therefore the Lower Steamboat Timber Sale is not likely to adversely affect 2012 critical habitat and will result in long term beneficial impacts by increasing structural diversity and heterogeneity to the stands being treated. The proposed thinning of NRF and dispersal habitat as part of the Lonely Oak Project will impact PCE’s 2, 3 and 4. There are a total of 93.4 acres of treatment proposed within critical habitat, and a total of 40 acres of NRF being downgraded to dispersal habitat within WCS-5 as a result of the implementation of the Lonely Oak project. The creation of a wetland within critical habitat is not consistent with the

61 restoration concepts of the final rule. This in conjunction with the downgrading of 40 acres of NRF the Lonely Oak Project may affect, and is likely to adversely affect the final critical habitat. However, long term, the thinning treatments proposed to release ponderosa pine and sugar pine, in conjunction with the application of prescribed fire within critical habitat will result in a more fire resilient landscape beneficial to maintain habitat suitability within this critical habitat subunit.

VI. Consistency with the Recovery Plan

In July of 2011 the revised recovery plan for the Northern Spotted Owl was finalized (USDI 2011). The Recovery Plan included four Recovery Criterion and 33 Recovery Actions, of which 4 Recovery Actions apply to the Lower Steamboat Timber Sale and the Lonely Oak Project.

Recovery Action 6: In moist forests managed for spotted owl habitat, land managers should implement silvicultural techniques in plantations, overstocked stands and modified younger stands to accelerate the development of structural complexity and biological diversity that will benefit spotted owl recovery. (p. III-19).

The Lower Steamboat Timber Sale is entirely within LSR and is designed to accelerate the development of structural complexity and biological diversity within the treated stands. Therefore the Lower Steamboat Timber Sale is entirely consistent with this recovery action. The Lonely Oak Project is not managing plantations, but is being designed to reduce stocking of conifers adjacent to ponderosa and sugar pines in the treatment units. Therefore the Lonely Oak Project is also consistent with this recovery action.

Recovery Action 10 - Conserve spotted owl sites and high value spotted owl habitat to provide additional demographic support to the spotted owl population.

Recovery Action 10 goes on to suggest that in unsurveyed owl habitat (like the both projects) the USFWS and Forest Service should work together to minimize impacts to potential spotted owl sites. All of the known and modeled owl home range and core use areas are at or above threshold levels for amounts of suitable habitat and will remain as such post treatment. Therefore the Lower Steamboat Timber Sale and Lonely Oak project are consistent with Recovery Action 10.

Recovery Action 32: Because spotted owl recovery requires well distributed, older and more structurally complex multi-layered conifer forests on Federal and non-federal lands across its range, land managers should work with the Service as described below to maintain and restore such habitat while allowing for other threats, such as fire and insects, to be addressed by restoration management actions. These high-quality spotted owl habitat stands are characterized as having large diameter trees, high amounts of canopy cover, and decadence components such as broken-topped live trees, mistletoe, cavities, large snags, and fallen trees.

There are no stands being proposed for treatment in either project that meet the criteria of Recovery Action 32. Therefore both projects are consistent with Recover Action 32.

VIII. Cumulative Effects

62

Section 7 of the ESA cumulative effects include the effects of future state, tribal, local or private actions that are reasonably certain to occur within the action area considered in this biological assessment. Future federal actions which are unrelated to the proposed projects are not considered in this section because they require separate consultation. No state-administered lands are located within the Umpqua National Forest. The Lonely Oak action area is entirely federal lands managed by the Umpqua National Forest. The Lower Steamboat action area does contain approximately 470 acres of private land and 666 acres of Roseburg District BLM. The private land is owned by Seneca Jones. Therefore it is reasonably certain that it will be clearcut within the near future. As this represents only 1% of the action area, there will be a minimal overall impact to the action area if any of this land is currently functioning as habitat and is subsequently removed. As the adjacent Forest Service and BLM lands are both LSR and critical habitat this loss of habitat would not result in an impact that would change the effects determination for the Lower Steamboat Timber Sale.

IX. Interrelated and Interdependent Actions

Interrelated actions are those that are part of a larger action and depend on the larger action for their justification. Interrelated actions are related to the federal action, but do not depend on the federal action. Interdependent actions are those that might occur independently of the larger action, but which have no independent utility apart from the action under consideration. Interdependent actions depend on the federal action and would make no sense without the federal action. The interdependent actions have been described in section VI as indirect effects.

X. Determinations

Table 19. Total impacts to Northern Spotted Owl dispersal and nesting, roosting and foraging habitat for both the Lower Steamboat Timber Sale and the Lonely Oak Project. NRF Habitat Dispersal Habitat Dispersal Habitat NRF Habitat Project Downgraded to Maintained Removed Maintained Dispersal Lower 802 392.4 0 0 Steamboat Lonely Oak 58 0 6 69

The Lower Steamboat Timber Sale will result in 802 acres of maintained dispersal habitat and 392.4 acres of dispersal habitat removed. As a result of the removal of the dispersal habitat for gap creation and temp road creation, along with the long term impacts of the silvicultural treatments to spotted owl prey habitat, the Lower Steamboat Timber Sale may affect, and is likely to adversely affect the spotted owl. . However, only 66.5 acres of dispersal habitat will be downgraded within critical habitat sub unit WCS-5, and no NRF habitat will be impacted. Therefore the Lower Steamboat Timber Sale is not likely to adversely affect 2012 critical habitat and will result in long term beneficial impacts by increasing structural diversity and heterogeneity to the stands being treated. The Lonely Oak project will maintain 58 acres of dispersal and 6 acres of NRF habitat, but will also downgrade 69 acres of NRF habitat to dispersal habitat. As a result of the downgrade in NRF habitat and the long term impacts of the silvicultural treatments to prey habitat, the Lonely Oak project may affect, and is likely to adversely affect the spotted owl. The creation of a wetland within critical habitat is not consistent with the restoration concepts of the final rule. This in conjunction with the

63 downgrading of 40 acres of NRF the Lonely Oak Project may affect, and is likely to adversely affect the final critical habitat. However, long term, the thinning treatments proposed to release ponderosa pine and sugar pine, in conjunction with the application of prescribed fire within critical habitat will result in a more fire resilient landscape beneficial to maintain habitat suitability within this critical habitat subunit.

XI. Literature Cited

Anderson, D.R. and K.P. Burnham. 1992. Demographic analysis of northern spotted owl populations. Pages 319–328 in Draft final recovery plan for the northern spotted owl. USDI Fish and Wildlife Service, Portland, Oregon.

Anthony, R.G., E.D. Forsman, A.B. Franklin, D.R. Anderson, K.P. Burnham, G.C. White, C.J. Schwarz, J. Nichols, J.E. Hines, G.S. Olson, S.H. Ackers, S. Andrews, B.L. Biswell, P.C. Carlson, L.V. Diller, K.M. Dugger, K.E. Fehring, T.L. Fleming, R.P. Gerhardt, S.A. Gremel, R.J. Gutiérrez, P.J. Happe, D.R. Herter, J.M. Higley, R.B. Horn, L.L. Irwin, P.J. Loschl, J.A. Reid, and S.G. Sovern. 2006. Status and trends in demography of northern spotted owls, 1985-2003. Status and trends in demography of northern spotted owls, 1985–2003. Wildlife Monograph No. 163.

Barrowclough, G.F. and R. J. Gutiérrez. 1990. Genetic variation and differentiation in the spotted owl. Auk 107:737-744.

Barrowclough, G.F., R.J. Gutiérrez, and J.G. Groth. 1999. Phylogeography of spotted owl (Strix occidentalis) populations based on mitochondrial DNA sequences; gene flow, genetic structure, and a novel biogeographic pattern. Evolution 53(3):919-931.

Barrowclough, G.F., J.G. Groth, and R.J. Gutiérrez. 2005. Genetic structure, introgression and a narrow hybrid zone between northern and California spotted owls (Strix occidentalis). Molecular Ecology 14:1109–1120.

Bart, J. 1995. Amount of suitable habitat and viability of northern spotted owls. Conservation Biology 9 (4):943-946.

Bart, J., and E.D. Forsman. 1992. Dependence of northern spotted owls Strix occidentalis caurina on old-growth forests in the western USA. Biological Conservation 62:95-100.

Bigley, R. and J. Franklin. 2004. Habitat trends. Chapter 6 in S. Courtney, editor. Scientific evaluation of the status of the northern spotted owl. Sustainable Ecosystems Institute, Portland, Oregon.

Bond, M.L., R.J. Gutiérrez, A.B. Franklin, W.S. LaHaye, C.A. May, and M.E. Seamans. 2002. Short-term effects of wildfires on spotted owl survival, site fidelity, mate fidelity, and reproductive success. Wildlife Society Bulletin 30:1022–1028.

Buchanan, J.B., L.L. Irwin, and E.L. McCutchen. 1995. Within-stand nest site selection by spotted owls in the eastern Washington Cascades. Journal of Wildlife Management 59:301-310.

Buchanan, J.B.,R. J. Gutiérrez, R.G. Anthony, T. Cullinan, L.V. Diller, E.D. Forsman, A.B. Franklin. 2007. A

64

synopsis of suggested approaches to address potential competitive interactions between Barred Owls (Strix varia) and Spotted Owls (S. occidentalis). Biological Invasions 9:679–69.

Bull, E.L, T.W. Heater, and A. Youngblood. 2004. Arboreal squirrel response to silvicultural treatment for dwarf mistletoe control in northeastern Oregon. Western Journal of Applied Forestry 19:133- 141

Burnham, K.P., D.R. Anderson, and G.C. White. 1994. Estimation of vital rates of the northern spotted owl. Colorado Cooperative Fish and Wildlife Research Unit, Colorado State University, Fort Collins, Colorado, USA.

Carey, A.B. and M.L. Johnson. 1995. Small mammals in managed, naturally young, and old-growth forests. Ecological Applications 5:336-352

Carey, A.B., Maguire, C.C., Biswell, B.L. and T. Wilson. 1999. Distribution and Abundance of Neotoma in Western Oregon and Washington. Northwest Science 73(2):65-80.

Carey, A.B. et al. 1999. Ecological Scale and Forest Development: Squirrels, Dietary Fungi, and Vascular Plants in Managed and Unmanaged Forests. Wildlife Monographs, No. 142. J. Wildl. Manage. Vol. 63, No. 1, January 1999.

Carey, A.B. 2000. Effects of new forest management strategies on squirrel populations. Ecological Applications 10: 248- 257.

Carey, A.B. and C.A. Harrington. 2001. Small mammals in young forests: implications for management for sustainability. Forest Ecology and Management 154:289-309

Carey, A.B. 2003. Biocomplexity and restoration of biodiversity in temperate coniferous forest: Inducing spatial heterogeneity with variable-density thinning. Forestry 76(2):127-136.

Carey, A.B. and M.L. Johnson. 1995. Small mammals in managed, naturally young, and old-growth forests. Ecological Applications 5:336-352.

Carroll, C., J. R. Dunk and A. Moilanen. 2010. Optimizing resiliency of reserve networks to climate change: multispecies conservation planning in the Pacific Northwest, USA. Global Change Biology 16:891-904.

Colgan, W., A.B. Carey, J.M. Trappe, R. Molina and D. Thysell. 1999. Diversity and productivity of hypogeous fungal sporocarps in a variably thinned Douglas-fir forest. Can. J. For. Res. 29(8): 1259–1268.

Collins, B.M.,S.L. Stephens, J.J. Moghaddas and J. Battles. 2010. Challenges and Approaches in Planning Fuel Treatments across Fire-Excluded Forested Landscapes. Journal of Forestry, Vol:108(1):24- 31.

Converse, S.J., G.C. White, K.L. Farris and S. Zack. 2006. Small Mammals and Forest Fuel Reduction: National-Scale Responses to Fire and Fire Surrogates. Ecological Applications, Vol:16(5): 1717- 1729.

65

Courtney, S.P, J.A. Blakesley, R.E. Bigley, M.L. Cody, J.P. Dumbacher, R.C. Fleischer, A.B. Franklin, J.F. Franklin, R.J. Gutiérrez, J.M. Marzluff and L. Sztukowski. 2004. Scientific evaluation of the status of the Northern Spotted Owl. Sustainable Ecosystems Institute Portland, Oregon.

Davis, Raymond J.; Dugger, Katie M.; Mohoric, Shawne; Evers, Louisa; Aney, William C. 2011. Northwest Forest Plan—the first 15 years (1994–2008): status and trends of northern spotted owl populations and habitats. Gen. Tech. Rep. PNWGTR- 850. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 147 p.

Delaney, D.K., T.G. Grubb, P. Beier, L.L. Pater, M. H. Reiser. 1999. Effects of Helicopter Noise on Mexican Spotted Owls. The Journal of Wildlife Management 63(1): 60-76.

Diller, L.V. and D.M. Thome. 1999. Population density of northern spotted owls in managed young- growth forests in coastal northern California. Journal of Raptor Research 33: 275–286.

Dugger, K.M., F. Wagner, R.G. Anthony, and G.S. Olson. 2005. The relationship between habitat characteristics and demographic performance of northern spotted owls in southern Oregon. The Condor 107:863-878.

Dunbar, D.L., B.P. Booth, E.D. Forsman, A.E. Hetherington, and D.J. Wilson. 1991. Status of the spotted owl, Strix occidentalis, and barred owl, Strix varia, in southwestern British Columbia. Canadian Field-Naturalist 105:464–468.

Folliard, L. 1993. Nest site characteristics of northern spotted owls in managed forest of northwest California. M.S. Thesis. University of Idaho, Moscow, Idaho.

Forsman, E.D. 1975. A preliminary investigation of the spotted owl in Oregon. M.S. thesis, Oregon State University, Corvallis. 127 pp.

Forsman, E.D., E.C. Meslow, and H.M. Wight. 1984. Distribution and biology of the spotted owl in Oregon. Wildlife Monographs 87:1-64.

Forsman, E.D., S. DeStefano, M.G. Raphael, and R.J. Gutierrez, (editors). 1996. Demography of the northern spotted owl. Studies in Avian Biology No. 17. Cooper Ornithology Society.

Forsman, E.D., I.A. Otto, S.G. Sovern, M. Taylor, D.W. Hays, H. Allen, S.L. Roberts, and D.E. Seaman. 2001. Spatial and temporal variation in diets of spotted owls in Washington. Journal of Raptor Research 35(2):141-150.

Forsman, E.D., R.G. Anthony, J.A. Reid, P.J. Loschl, S.G. Sovern, M. Taylor, B.L. Biswell, A. Ellingson, E.C. Meslow, G.S. Miller, K.A. Swindle, J.A. Thrailkill, F.F. Wagner, and D. E. Seaman. 2002. Natal and breeding dispersal of northern spotted owls. Wildlife Monographs 149:1-35.

Forsman, E.D., R.G. Anthony, E.C. Meslow, and C.J. Zabel. 2004. Diets and foraging behavior of northern spotted owls in Oregon. Journal of Raptor Research 38(3):214-230.

66

Forsman, E.D., T.J. Kaminski, J.C. Lewis, K.J. Maurice, S.G. Sovern, C. Ferland, and E.M. Glenn. 2005. Home range and habitat use of northern spotted owls on the Olympic Peninsula, Washington. Journal of Raptor Research 39(4):365-377.

Forsman, E. D., R. G. Anthony, K. M. Dugger, E. M. Glenn, A. B. Franklin, G. C. White, C. J. Schwarz, K. P. Burnham, D. R. Anderson, J. D. Nichols, J. E. Hines, J. B. Lint, R. J. Davis, S. H. Ackers, L. S. Andrews, B. L. Biswell, P. C. Carlson, L. V. Diller, S. A.Gremel, D. R. Herter, J. M. Higley, R. B. Horn, J. A. Reid, J. Rockweit, J. Schaberl, T. J. Snetsinger, and S. G. Sovern. 2011. Population demography of Northern Spotted Owls. Studies in Avian Biology 40.

Franklin, A.B. 1992. Population regulation in northern spotted owls: theoretical implications for management. Pp 815-827 in D. R. McCullough and R. H. Barrett (eds.)., Wildlife 2001: populations. Elsevier Applied Sciences, London, England.

Franklin, A.B., D.R. Anderson, R.J. Gutiérrez, and K.P. Burnham. 2000. Climate, habitat quality, and fitness in northern spotted owl populations in northwestern California. Ecological Monographs, 70(4): 539-590.

Gaines, W.L., R.A. Strand, and S.D. Piper. 1997. Effects of the Hatchery Complex Fires on northern spotted owls in the eastern Washington Cascades. Pages 123-129 in Dr. J.M. Greenlee, ed. Proceedings of the First Conference on Fire Effects on Rare and Endangered Species and Habitats, November 13-16, 1995. International Association of Wildland Fire. Coeur d’Alene, ID.

Gremel, S. 2005. Factors controlling distribution and demography of northern spotted owls in a reserved landscape. Thesis, University of Washington, Seattle.

Gomez, D. M., R. G. Anthony, J. P. Hayes, and Martin. 2005. Influence Of Thinning Of Douglas-Fir Forests On Population Parameters And Diet Of Northern Flying Squirrels. Journal of Wildlife Management 69:1670-1682.

Gutiérrez, R.J. 1996. Biology and distribution of the northern spotted owl. Pages 2-5 in E.D. Forsman et al., editors. Demography of the northern spotted owl. Studies in Avian Biology No. 17. Cooper Ornithological Society.

Gutiérrez, R.J., A.B. Franklin, and W.S. LaHaye. 1995. Spotted owl (Strix occidentalis) in: A. Poole and F. Gill, editors. The birds of North America, No. 179. The Academy of Natural Sciences and The American Ornithologists' Union, Washington, D.C. 28 pages.

Haig, S.M., T.D. Mullins, and E.D. Forsman. 2004. Subspecific relationships and genetic structure in the spotted owl. Conservation Genetics 5:683–705.

Hamer, T.E., S.G. Seim, and K.R. Dixon. 1989. Northern spotted owl and northern barred owl habitat use and home range size in Washington: preliminary report. Washington Department of Wildlife, Olympia, Washington.

Hamer, T.E., E.D. Forsman, A.D. Fuchs, and M.L. Walters. 1994. Hybridization between barred and spotted owls. Auk 111(2):487-492.

67

Hamer, T.E., D.L. Hays, C.M. Senger, and E.D. Forsman. 2001. Diets of northern barred owls and northern spotted owls in an area of sympatry. Journal of Raptor Research 35(3):221-227.

Hamer, T. E., E. D. Forsman, and E. M. Glenn. 2009. Home range attributes and habitat selection of barred owls and spotted owls in an area of sympatry. The Condor 109:750-768.

Healey, S.P., W.B. Cohen, T.A. Spies, M. Moeur, D. Pflugmacher, M.G. Whitley and M. Lefsky. 2008. The Relative Impact of Harvest and Fire upon Landscape-Level Dynamics of Older Forests: Lessons from the Northwest Forest Plan. Ecosystems (2008) 11: 1106–1119

Hebers, J., and W. Klenner. 2007. Effects of logging pattern and intensity on squirrel demography. Journal of Wildlife Management 69:1670-1682.

Hershey, K.T., E.C. Meslow, and F.L. Ramsey. 1998. Characteristics of forests at spotted owl nest sites in the Pacific Northwest. Journal of Wildlife Management 62(4):1398-1410.

Herter, D.R. and L.L. Hicks. 2000. Barred owl and spotted owl populations and habitat in the central Cascade Range of Washington. Journal of Raptor Research 34:279–286.

Hoberg, E.P., G.S. Miller, E. Wallner-Pendleton, and O.R. Hedstrom. 1989. Helminth parasites of northern spotted owls (Strix occidentalis caurina). Journal of Wildlife Diseases 25:246–251.

Irwin, L.L., D.F. Rock, and G.P. Miller. 2000. Stand structures used by northern spotted owls in managed forests. Journal of Raptor Research 34:175–186.

Iverson, W.F. 1993. Is the barred owl displacing the spotted owl in western Washington? M.S. Thesis, Western Washington University, Bellingham, Washington.

Iverson, W.F. 2004. Reproductive success of spotted owls sympatric with barred owls in western Washington. Journal of Raptor Research 38(1):88-91.

Johnson, K. N., J. F. Franklin, J. W. Thomas, and J. Gordon. 1991. Alternatives for management of late successional forests of the Pacific Northwest. A report to the Agriculture and Merchant and Marine Fisheries Committees of the U.S. House of Representative. Scientific Panel on Late Successional Forest Ecosystems, Washington, D.C.

Kelly, E.G., E.D. Forsman, and R.G. Anthony. 2003. Are barred owls replacing spotted owls? The Condor 105:45-53.

Kelly, E.G. and E.D. Forsman. 2004. Recent records of hybridization between barred owls (Strix varia) and northern spotted owls (S. occidentalis caurina). Auk 121:806-810.

King, G.M., K.R. Bevis, M.A. Rowe, and E.E. Haspotted owln. 1997. Spotted owls use of habitat impacted by 1994 fires on the Yakama Indian Reservation: three years post fire. Journal of Forestry 95:21–25.

68

Knight, R.L., and S.K. Skagen. 1998. Effects of recreational disturbance on birds of prey: a review. Pages 355-359 in R.L. Glinski et al., editors. Proceedings of the Southwest Raptor Management Symposium and Workshop, National Wildlife Federation, Washington, D.C.

LaHaye, W.S., R.J. Guiterrez, and J.R. Dunk. 2001. Natal dispersion of the spotted owl in southern California: dispersal profile of an insular population. The Condor 103:691-700.

Laidig, K.J., and D.S. Dobkin. 1995. Spatial overlap and habitat association of Barred Owls and Great Horned Owls in southern New Jersey. Journal of Raptor Research 29:151–157.

Lee, D.C. and L. Irwin. 2005. Assessing risks to spotted owls from forest thinning in fire-adapted forests of the western United States. Forest Ecology and Management 211:191-209.

Lehmkuhl, J.F., K.D. Kistler, J.S. Begley, and J. Boulanger. 2006. Demography of Northern Flying Squirrels Informs Ecosystem Management of Western Interior Forests. Ecological Applications, Vol:16(2):584-600

Lenihan, J.M., R. Drapek, D. Bachelet and R.P. Neilson. 2003. Climate change effects on vegetation distribution, carbon, and fire in California. Ecological Applications 13(6)1667-1681.

Littell, J.S., D. McKenzie, D.L. Peterson, A.L. Westerling. 2009. Climate and wildfire area burned in western U.S ecoprovinces, 1916–2003. Ecological Applications 2009 19:4, 1003-1021

Loehle, C. 2004. Applying landscape principles to fire hazard reduction. Forest Ecology and Management. 198: 261-267.

Manning, T., J.C. Hagar and B.C. McComb. 2012. Thinning of young Douglas-fir forests decreases density of northern flying squirrels in the Oregon Cascades. Forest Ecology and Management 264:115–124.

McComb, B., R. Pabst, T. Manning and C.A. Jordan. 2009. Dead Wood and Small Mammal Reponses to Three Alternative Thinning Strategies in Young Douglas-fir Forests. Unpublished report.

McGarigal, K., R.G. Anthony, and F.B. Issacs. 1991. Interactions of humans and bald eagles on the Columbia River estuary. Wildlife Monographs 115:1-47.

Meyer, M.D., M.P. North and D.A. Kelt. 2005. Short-term effects of fire and forest thinning on truffle abundance and consumption by Neotamias speciosus in the Sierra Nevada of California. Can. J. For. Res. 35: 1061–1070.

Millar, C.I., N.L. Stephen and S.L. Stephens. 2006. Climate Change and Forests of the Future: Managing in the Face of Uncertainty. Ecological Applications, 17(8): 2145–2151.

Miller, G.S. 1989. Dispersal of juvenile spotted owls in western Oregon. Thesis, Oregon State University, Corvallis.

Moen, C.A., A.B. Franklin, and R.J. Gutiérrez. 1991. Age determination of subadult northern spotted owls in northwest California. Wildlife Society Bulletin 19:489–493.

69

Moeur, M., T.A. Spies, M. Henstrom, J.R. Martin, J. Alegria, J. Browning, J. Cissel, W.B. Cohen, T.E. Demeo, S. Healey, R. Warbington. 2005. Northwest Forest Plan–The first 10 years (1994-2003): status and trend of late-successional and old-growth forest. Gen. Tech. Rep. PNW-GTR-646. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 142 p.

Muir, P.S., R.L. Mattingly, J.C. Tappeiner II, J.D. Bailey, W.E. Elliott, J.C. Hagar, J.C. Miller, E.B. Peterson, and E.E. Starkey. 2002. Managing for biodiversity in young Douglas-fir forests of western Oregon. U.S. Geological Survey, Biological Resources Division, Biological Science Report USGS/BRD/BSR–2002-0006. 76 pp.

Noon, B.R. and J.A. Blakesley. 2006. Conservation of the northern spotted owl under the Northwest Forest Plan. Conservation Biology 20:288–296.

Olson, G.S., R.G. Anthony, E.D. Forsman, S.H. Ackers, P.J. Loschl, J.A. Reid, K.M Dugger, E.M. Glenn, and W.J. Ripple. 2005. Modeling of site occupancy dynamics for northern spotted owls, with emphasis on the effects of barred owls. Journal of Wildlife Management 69(3):918-932.

Olson, G.S., E.M. Glenn, R.G. Anthony, E.D. Forsman, J.A. Reid, P.J. Loschl, and W.J. Ripple. 2004. Modeling demographic performance of northern spotted owls relative to forest habitat in Oregon. Journal of Wildlife Management 68:1039–1053.

Paton, P.W.C., C.J. Zabel, D.L. Neal, G.N. Steger, N.G. Tilghman, B.R. Noon. 1991. Effects of Radio Tags on Spotted Owls. The Journal of Wildlife Management 55(4): 617-622

Parmesan, C. and G. Yohe. 2003. A globally coherent fingerprint of climate change impacts across natural systems. Nature 421:37-42.

Pearson, Robert R. and K.B. Livezey. 2003. Distribution, numbers, and site characteristics of spotted owls and barred owls in the Cascade Mountains of Washington. J. Raptor Res. 37(4):265-275.

Ransome D.B. and T.P. Sullivan. 2002. Short-term population dynamics of Glaucomys sabrinus and Tamiasciurus douglasii in commercially thinned and unthinned stands of coastal coniferous forest. Can. J. For. Res. 32(11): 2043–2050.

Rosenberg, D.K. and R.G. Anthony. 1992. Characteristics of northern flying squirrel populations in young second- and old-growth forests in western Oregon. Can. J. Zool. 70:161-166.

Rosenberg, D.K., K.A. Swindle, and R.G. Anthony. 2003. Influence of prey abundance on northern spotted owl reproductive success in western Oregon. Can. J. Zool. 81: 1715–1725.

Sakai, H.F.; Noon, B.R. 1993. Dusky-footed woodrat abundance in different-aged forests in northwestern California Journal of Wildlife Management 57(2):373-382.

Sapolsky, R. M., L. M. Romero, and A. U. Munck. 2000. How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews 21:55-89.

70

Schwartz, M.W., L.R. Iverson, A.M. Prasad, S.N. Matthews and R.J. O’connor. 2006. Predicting Extinctions As A Result Of Climate Change. Ecology, 87(7): 1611–1615.

Schmidt, K. 2006. Northern spotted owl monitoring and inventory, Redwood National and State Parks, 2005 annual report. Redwood National and State Parks, Orick, California.

Sisco, C.L. 1990. Seasonal home range and habitat ecology of spotted owls in northwestern California. M.S. Thesis. Humboldt State University, Arcata, California.

Singleton, P. H., J. F. Lehmkuhl, W. L. Gaines, and S. A. Graham. 2010. Barred Owl Space Use and Habitat Selection in the Eastern Cascades, Washington. Journal of Wildlife Management 74:285-294.

Solis, D.M., and R.J. Gutiérrez. 1990. Summer habitat ecology of northern spotted owls in northwestern California. The Condor 92:739-748.

Sovern, S.G., E.D. Forsman, B.L. Biswell, D.N. Rolph, and M. Taylor. 1994. Diurnal behavior of the spotted owl in Washington. Condor 96:200–202.

Spies, T.A., J.D. Miller, J.B. Buchanan, J.F. Lehmkuhl, J.F. Franklin, S.P. Healey, P.F. Hessburg, H.D. Safford, W.B. Cohen, R.S.H. Kennedy, E.E. Knapp, J.K. Agee and M. Moeur. 2010. Underestimating Risks to the Northern Spotted Owl in Fire-Prone Forests: Response to Haspotted owln et al. Conservation Biology Vol:24(1): 330–333.

Stephens,S.L., J.J. Moghaddas, C. Edminster, C.E. Fiedler, S. Haase, M. Harrington, J.E. Keeley, E.E. Knapp, J.D. Mciver, K. Metlen, C.N. Skinner, and A.Youngblood. 2009. Fire treatment effects on vegetation structure, fuels, and potential fire severity in western U.S. forests. Ecological Applications, 19(2)305–320.

Suzuki, N. and J.P. Hayes. 2003. Effects of Thinning on Small Mammals in Oregon Coastal Forests. The Journal of Wildlife Management Vol:67(2): 352-371

Tappeiner et al, 1997. Density, ages and growth rates in old-growth and young-growth forests in coastal Oregon. Can. J. For. Res. 27:638-648).

Tempel, D. J. and R. J. Gutiérrez. 2003. Fecal corticosterone levels in California spotted owls exposed to low-intensity chainsaw noise. Wildlife Society Bulletin 31:698-702.

Tempel, D. J. and R. J. Gutiérrez. 2004. Factors related to fecal corticosterone levels in California spotted owls: implications for assessing chronic stress. Conservation Biology 18:538-547.

Thomas, J.W., E.D. Forsman, J.B. Lint, E.C. Meslow, B.R. Noon, and J. Verner. 1990. A conservation strategy for the northern spotted owl. A report by the Interagency Scientific Committee to address the conservation of the northern spotted owl. USDA, Forest Service, and U. S. Department of the Interior, Fish and Wildlife Service, Bureau of Land Management, and National Park Service. Portland, OR.

71

Thomas, J.W., M.G. Raphael, R.G. Anthony, E.D. Forsman, A.G. Gunderson, R.S. Holthausen, B.G. Marcot, G.H. Reeves, J.R. Sedell, and D.M. Solis. 1993. Viability assessments and management considerations for species associated with late-successional and old-growth forests of the Pacific Northwest. USDA Forest Service, Portland, Oregon.

Thome, D.M., C.J. Zabel and L.V. Diller. 1999. Forest stand characteristics and reproduction of northern spotted owls in managed north-coastal California forests, Journal of Wildlife Management 63:44– 59.

United States Department of Agriculture, US Forest Service. 1990. Land and resource management plan, Umpqua National Forest. Pacific Northwest Region. http://www.fs.fed.us/r6/umpqua/projects/plandocs/umpqua-lrmp.pdf

USDA Forest Service. 2008. Field Guide for Danger Tree Identification and Response.

USDI Fish and Wildlife Service, USDI Bureau of Land Management and USDA Forest Service. 2008. Methodology for estimating the number of northern spotted owls affected by proposed federal actions. Version 2.0. Oregon Fish and Wildlife Office, Fish and Wildlife Service, Portland, OR.

USDI. U.S. Fish and Wildlife Service. 1983a. Endangered and threatened species listing and recovery priority guidelines. Federal Register 48:43098–43105. ------. 1983b. Endangered and threatened species listing and recovery priority guidelines: correction. Federal Register 48:51985. ------. 1989. The northern spotted owl; a status review supplement. US Fish and Wildlife Service, Region 1, Portland, OR

------. 1990b. Endangered and threatened wildlife and plants; determination of threatened status for the northern spotted owl. Federal Register 55:26114–26194.

------. 1992. Endangered and Threatened Wildlife and Plants; Determination of Critical Habitat for the Northern Spotted Owl. January 15, 1992. Federal Register: Vol: 57(10):1796-1797.

------. 1992b. Draft final recovery plan for the northern spotted owl. U.S. Fish and Wildlife Service, Portland, Oregon.

------. 1994. Letter from D.C. Frederick (USFWS) to J. Belcher (Washington Department of Natural Resources) dated February 1, 1994, regarding correction to spotted owl home-range data for the Olympic Peninsula. U.S. Fish and Wildlife Service, Olympia, Washington.

------. 2004a. Northern Spotted Owl: Five Year Review Summary and Evaluation. USDI Fish and Wildlife Service, Portland, Oregon. 73 pages.

------. 2008b. Endangered and threatened wildlife and plants; revised designation of critical habitat for the northern spotted owl; final rule. 50 CFR Part 17, Federal Register, August 13, 2008. Pages 47326- 47522.

------. 2011. Revised Recovery Plan for the Northern Spotted Owl, Strix occidentalis caurina.

72

U.S. Fish and Wildlife Service, Portland, Oregon.

------. 2012. Endangered and Threatened Wildlife and Plants; Designation of Revised Critical Habitat for the Northern Spotted Owl; Final Rule. 50 CFR Part 17, Federal Register / Vol. 77, No. 233 / December 4, 2012. Pages 71876-72068.

Van Lanen, N.J.. 2010. Out with the old and in with the new? Investigating competition between barred owls (Strix varia) and Northern spotted owls (Strix occidentalis caurina) in Northwestern California with a playback experiment. Unpublished Thesis, March 24, 2010, Colorado State University, Fort Collins, CO. 133 pp.

Ward, J. W. Jr. 1990. Spotted owl reproduction, diet and prey abundance in northwest California. M.S. Thesis. Humboldt State University, Arcata, California.

Ward, J. W. Jr., R.J. Gutiérrez, and B.R. Noon. 1998. Habitat selection by northern spotted owls: the consequences of prey selection and distribution. The Condor 100:79-92.

Wasser, S. K., K. Bevis, G. King, and E. Haspotted owln. 1997. Noninvasive physiological measures of disturbance in the Northern Spotted Owl. Conservation Biology 11:1019-1022.

Weathers, W.W., P.J. Hodum and J.A. Blakesley. 2001. Thermal ecology and ecological energetics of the California spotted owl. The Condor 103(4):678.

Westerling, A. L., H. G. Hidalgo, D. R. Cayan, and T. W. Swetnam. 2006. Warming and Earlier Spring Increase Western U.S. Forest Wildfire Activity. Science 313:940-943 http://www.sciencemag.org/cgi/content/abstract/313/5789/5940.

Wilson, T.M. 2010. Limiting factors for northern flying squirrels (Glaucomys sabrinus) in the Pacific Northwest: A spatio-temporal analysis. In Fullfillment of PhD for Union Institute and University, 2010, 220 pages.

Wilson, T.M. 2011. Email communication with author. 10/21/2011.

Wilson, Todd M.; Forsman, Eric D. 2013. Thinning effects on spotted owl prey and other forest-dwelling small mammals. In: Anderson, Paul D.; Ronnenberg, Kathryn L., eds. Density management for the 21st Century: west side story. Gen.Tech. Rep. PNW-GTR-880. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station: 79–90.

Zabel, C. J., K.M. McKelvey, and J.P. Ward, Jr. 1995. Influence of primary prey on home-range size and habitat-use patterns of northern spotted owls (Strix occidentalis caurina). Canadian Journal of Zoology 73:433-439.

Zabel, C. J., J.R. Dunk, H.B. Stauffer, L.M. Roberts, B.S. Mulder, and A. Wright. 2003. Northern spotted owl habitat models for research and management application in California (USA). Ecological Applications 13:1027–1040.

73

Appendix A. Project Design Features for the Lower Steamboat Timber Sale

1. Retain existing large down wood (>6 inch diameter) and snags (>9 inch dbh) to the extent practical and safe. Where feasible, avoid mechanical impacts and movement of large down wood and leave felled snags on site. Care will be taken when yarding to attempt to avoid loss of bark on downed wood. If possible, directionally fall and yard trees away from large downed wood.

2. If feasible, in skyline units, retain all trees used as anchors in the skyline operation as long as they do not pose a hazard.

3. OSHA requires that hazardous trees/snags be felled to protect workers on the ground during forest operations. Snags that must be felled for safety reasons should be retained to help attain down wood requirements.

4. When felling hazard trees retain as high of a stump as is operationally safe to do so.

5. All trees damaged during harvest operation, such as intermediate support trees or line damage trees, would be retained to mitigate the decreased rate of snag recruitment caused by thinning and harvest activities.

6. Leave all guyline tail-hold trees outside of unit boundaries.

7. To protect nesting spotted owls, for proposed and connected actions that create above-ambient noise levels (i.e. road maintenance, brushing, subsoiling, etc.), abide by the terms and conditions in the programmatic Biological Opinion (FWS-1-15-03-F-0454). When possible, do not schedule these activities to occur between March 1 and July 15.

8. Anchor Tree Selection: Cable logging systems may use healthy green trees to anchor rigging (tailholds) and yarders (guyline trees). Anchor tree removal will not occur in any habitat type. The smallest possible anchor trees will be selected in all instances, trees with suitable spotted owl nest structures will be avoided when possible, and anchor trees will be left standing when feasible. These measures will help to reduce impacts to spotted owl habitat features. Anchor Tree felling of occupied nest trees will be strongly avoided where- and whenever possible (due to lack of current surveys throughout the planning areas, the Forest cannot be absolutely certain no occupied nest trees will be felled). The following features will minimize the likelihood of felling occupied nest trees:

1) when large guyline trees are necessary, trees closer to the road will be selected in lieu of trees farther into the adjacent stand; 2) trees with characteristics suitable for spotted owl nesting will be avoided wherever possible; 3) use of guyline trees necessitating felling of large snags in the adjacent stands will be avoided wherever possible; 4) use of guyline trees close enough to possible nest trees that felling them will disrupt the current micro-climatic conditions associated with said possible nest trees will be avoided wherever possible.

74

5) guyline trees felled for safety reasons will be left on site when felled outside of units after the cessation of logging operations. All anchor trees outside of unit boundaries are to be retained as either live trees or snags if they have been topped.

75