Hudec et al. Front Matter 1

1 Climate Change Vulnerability and Adaptation in

2 Southwest 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Editors 33 J.L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot 34 National Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650; J.E. 35 Halofsky is a research ecologist, University of Washington, College of the Environment, School 36 of Environmental and Forest Sciences, Box 352100, , WA 98195-2100; D.L. Peterson is 37 a senior research biological scientist, U.S. Department of Agriculture, Forest Service, Pacific 38 Northwest Research Station, 400 N 34th St., Suite 201, Seattle, WA 98103; J.J. Ho is a research 39 economist, University of Washington, College of the Environment, School of Environmental and 40 Forest Sciences,DRAFT Box 352100, Seattle, WA 98195-2100. Hudec et al. Front Matter 2

41 Climate Change Vulnerability and Adaptation in

42 Southwest Washington 43 44 J.L. Hudec, J.E. Halofsky, D.L. Peterson, and J.J. Ho 45 46 47 Editors 48 49 50 51 52 53 U.S. Department of Agriculture, Forest Service 54 Pacific Northwest Research Station 55 Portland, 56 General Technical Report PNW-GTR-xxxx 57 Month year 58

DRAFT Hudec et al. Front Matter 3

59 Abstract 60 61 Hudec, J.L.; Halofsky, J.E.; Peterson, D.L.; Ho, J.J., eds. 201X. Climate change vulnerability 62 and adaptation in Southwest Washington. Gen. Tech. Rep. PNW-GTR-xxx. Portland, OR: 63 U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. Xxx p. 64 65 The Southwest Washington Adaptation Partnership (SWAP) was developed to identify climate 66 change issues relevant for resource management in Southwest Washington, specifically on 67 Gifford Pinchot National Forest. This science-management partnership assessed the vulnerability 68 of natural resources to climate change, and developed adaptation options that minimize negative 69 impacts of climate change on resources of concern and facilitate transition of diverse ecosystems 70 to a warmer climate. The vulnerability assessment focuses on fish and aquatic habitat, 71 vegetation, special habitats, recreation, and ecosystem services. 72 Projected changes in climate and hydrology will have far-reaching effects on aquatic and 73 terrestrial ecosystems, especially as frequency of extreme climatic events (drought, low 74 snowpack) and ecological disturbances (flooding, wildfire, insect outbreaks) increases. 75 Distribution and abundance of cold-water fish species are expected to decrease in response to 76 higher water temperature, although effects will vary as a function of local habitat and 77 competition with nonnative fish. 78 Higher air temperature, through its influence on soil moisture, is expected to cause 79 gradual changes in the distribution and abundance of plant species, with drought-tolerant species 80 becoming more dominant. Increased frequency and extent of wildfire will facilitate vegetation 81 change, in some cases leading to altered structure and function of ecosystems (e.g., more forest 82 area in younger age classes). Special habitats such as riparian areas and wetlands are expected to 83 be especially sensitive to altered soil moisture, especially as drought frequency increases. 84 The effects of climate change on recreation activities are difficult to project, although 85 warmer temperatures are expected to create more opportunities for warm-weather activities (e.g., 86 hiking, camping) and fewer opportunities for snow-based activities (e.g., skiing, snowmobiling). 87 Recreationists modify their activities according to current conditions, but recreation management 88 by federal agencies has generally not been so flexible. 89 Of the ecosystem services considered in the assessment, timber supply and carbon 90 sequestration may be affected by increasing frequency and extent of disturbances. Native 91 pollinators may be affected by altered vegetation distribution and phenological mismatches 92 between insects and plants. 93 Resource managers convened at a SWAP workshop and developed adaptation options in 94 response to the vulnerabilities identified in each resource area, including both high-level 95 strategies and on-the-ground tactics. Many adaptation options are intended to increase the 96 resilience of aquatic and terrestrial ecosystems, or to reduce the effects of existing stressors (e.g., 97 removal of nonnative species). In terrestrial systems, a dominant theme of adaptation in 98 Southwest Washington is to accelerate restoration and fuel treatments, particularly in drier forest 99 types, to reduce the undesirable effects of extreme events and high-severity disturbances 100 (wildfire, insects). In aquatic systems, a dominant theme is to restore the structure and function 101 of streams to retain cold water for fish and other aquatic organisms. Many adaptation options can 102 accomplish multiple outcomes; for example, fuel treatments in dry forests reduce fire intensity, 103 which in turn reduces erosion that would degrade water quality and fish habitat. Furthermore, 104 many existingDRAFT management practices are already “climate-informed” or require minor adjustment Hudec et al. Front Matter 4

105 to make them so. Long-term monitoring is needed to detect climate change effects on natural 106 resources of concern and to evaluate the effectiveness of adaptation options that are 107 implemented. 108 109 Keywords: Adaptation, aquatic ecosystems, climate change, fire, climate-informed 110 management, ecosystem services, fisheries, recreation, science-management partnership, 111 Southwest Washington, terrestrial ecosystems, vegetation. 112 113 114

DRAFT Hudec et al. Front Matter 5

115 Summary 116 117 The Southwest Washington Adaptation Partnership (SWAP) is a science-management 118 partnership consisting of Gifford Pinchot National Forest, the U.S. Forest Service Pacific 119 Northwest Region and Pacific Northwest Research Station, and the University of Washington. 120 These organizations worked together over a period of two years to identify climate change issues 121 relevant to resource management in Southwest Washington and to find solutions that can 122 minimize undesirable effects of climate change and facilitate transition of diverse ecosystems to 123 a warmer climate. SWAP provided education, conducted a climate change vulnerability 124 assessment, and developed adaptation options for Gifford Pinchot National Forest and adjacent 125 landowners. 126 Global climate models for a high-end greenhouse gas emission scenario (RCP 8.5; 127 comparable to current emissions) project that climatic warming will continue throughout the 21st 128 century. Compared to observed historical temperature (1950-1979), average warming is 129 projected to increase from 1.3 to 2.3 °C from 2010 to 2039, 2.5 to 4.2 °C for 2040-2069, and 4.3 130 to 6.4 °C for 2070-2099. Seasonally, the largest increases in temperature are projected for 131 summer (+5.7-6.8 °C on average for June–August in 2070-2099). Mean summer precipitation is 132 projected to decrease from 162 mm historically to 87-21 mm by the end of the century, while 133 extreme precipitation events are likely to increase. 134 Projected changes in climate and hydrology will have far-reaching effects on aquatic and 135 terrestrial ecosystems, especially as frequency of extreme climatic events (drought, low 136 snowpack) and associated effects on ecological disturbance (flooding, wildfire, insect outbreaks) 137 increase. Vulnerability assessment and development of adaptation options for Southwest 138 Washington conclude the following: 139 140 Fisheries and aquatic habitat— 141 • Effects: Higher winter streamflows, decreased summer streamflows, and warmer water 142 temperature will reduce habitat quality and extent for cold-water-dependent fish species. 143 Anadromous fish species will be susceptible to higher thermal stress during summer 144 upstream migrations, reduced access to upstream spawning areas, increased pre-spawn 145 mortality rates or reduced viability of eggs and embryos, and harsher conditions due to 146 greater disturbances while juveniles rear in streams or migrate downstream. Based on 147 projected stream temperatures in a warmer climate, coho salmon, which currently access 584 148 km of streams and rivers, may lose 40-65 percent of their current habitat. Chinook salmon 149 occupy a more restricted area (approximately 359 km) within the same river systems as coho, 150 and are susceptible to the same amount of habitat loss. Steelhead occupy the largest extent 151 (900 km of streams) of the anadromous fish species. Bull trout, which require very cold 152 water, live in fragmented habitats and have relatively small populations. In addition to 153 aforementioned stressors, bull trout will be increasingly susceptible to competition with 154 brook trout and other nonnative species in a warmer climate. 155 • Adaptation options: Primary adaptation strategies for fisheries and aquatic habitat focus on 156 storing more water on the landscape, increasing resilience to disturbance, maintaining and 157 restoring riparian and wetland vegetation complexity, and maintaining and restoring natural 158 thermal conditions in streams. Strategies in response to increased peak streamflows include 159 increasing spawning habitat resilience by restoring stream structure and processes and by 160 reducingDRAFT threats from roads and infrastructure, particularly in floodplains. A key strategy to Hudec et al. Front Matter 6

161 minimize the negative impacts of lower summer streamflows on habitat quality is to decrease 162 fragmentation of the stream network so fish can access suitable habitat. Restoring and 163 maintaining habitat quality and protecting cold-water refugia will help to mitigate effects of 164 increased stream temperatures. To reduce post-fire stream sedimentation, forest thinning and 165 prescribed fire can be used proactively to reduce fire severity and extent in dry forests. 166 167 Vegetation— 168 • Effects: Higher air temperature, through its influence on soil moisture, is expected to cause 169 gradual changes in the abundance and distribution of vegetation species, with more drought 170 tolerant species being more competitive. Ecological disturbance, mostly through increased 171 occurrence of wildfire and insect outbreaks, will be the primary facilitator of vegetation 172 change, and future forest landscapes may be dominated by younger age classes of trees. 173 Projections generally show vegetation zones shifting from their current positions to higher 174 elevations. 175 Alpine zone: Alpine vegetation is expected to be sensitive to changes in climate because of 176 potential for altered hydrologic regimes, limited reproductive capacity of some species, 177 isolation, and limited adaptive capacity. Short growing season, poor soil conditions, and 178 frequent disturbance hinder reproductive success of many alpine species. Lower snowpack 179 may lead to increased growth and productivity in the short term, but competitors from lower 180 elevations are expected to move to higher elevations. Isolated and endemic populations have 181 lower adaptive capacity and a higher risk of extinction than species with range continuity. 182 Parkland zone: Subalpine parklands are commonly associated with mountain hemlock, 183 subalpine fir, and whitebark pine. Closed-canopy forests have been increasing at the low- 184 elevation end of parklands through increased conifer establishment over the past century. The 185 parkland zone may move upward in elevation as higher temperatures and changes in timing 186 and amount of snowpack continue to favor conifer establishment and forest-dominated 187 systems at higher elevations. 188 Mountain hemlock zone: The mountain hemlock zone is projected to gradually contract in 189 area through the mid- to late-21st century. Earlier spring snowmelt could result in a longer 190 summer dry period, and area burned by high-severity fires may increase. Research suggests 191 that mountain hemlock tree growth near treeline could increase as the energy limitation of 192 this species is alleviated in a warmer climate. However, growth of mountain hemlock trees at 193 lower elevations may decrease where growth is limited by low soil moisture in summer. 194 Mountain hemlock zone may also be affected by more disturbance, and warmer temperatures 195 may favor the establishment of lower elevation species following disturbance. 196 Subalpine fir zone: Warming temperatures and decreased snowpack are expected to increase 197 subalpine fir tree growth near treeline. However, subalpine fir tree growth and seedling 198 establishment could decrease at low-elevation sites in the zone in response to drought stress. 199 Warmer temperatures may favor the establishment of grand fir trees following disturbance 200 and result in a shift in the grand fir zone to higher elevations currently occupied by subalpine 201 fir zone. 202 Pacific silver fir zone: A warmer climate could favor a transition of Pacific silver fir to higher 203 elevations, possibly reducing the dominance of mountain hemlock in some locations. Areas 204 of Pacific silver fir zone that are currently continuous may become disjunct as the species 205 movesDRAFT up in elevation toward isolated mountain peaks. An increase in area burned would Hudec et al. Front Matter 7

206 result in a larger portion of the zone in the grass-forb and post-disturbance structural stage 207 and a smaller portion in the large diameter, multi-story stage. 208 Western hemlock zone: A warmer climate with drier summers could favor a transition of 209 western hemlock zone to westside Douglas-fir zone. The western hemlock zone, in turn, may 210 move up in elevation, displacing the current lower extent of the Pacific silver fir zone. The 211 western hemlock zone is expected to remain continuous for genetic mobility and migration as 212 the zone moves up in elevation. 213 Eastside Douglas-fir zone: Eastside Douglas-fir zone will likely expand westward and into 214 areas currently occupied by the grand fir zone. The driest sites in the eastside Douglas-fir 215 zone could become increasingly dominated by ponderosa pine, especially if drought and 216 wildfire frequency increase in the future. 217 Westside Douglas-fir zone: Fire is the most significant natural disturbance in the westside 218 Douglas-fir zone, but windstorms, insects, and pathogens also affect the area. Warmer 219 climate with drier summers and more area burned would favor expansion of the westside 220 Douglas-fir zone into some portions of the western hemlock zone. An increase in area burned 221 may also facilitate increased dominance of grasses, forbs, and early-seral vegetation 222 structure. 223 Grand fir zone: Grand fir is expected to expand into drier portions of the western hemlock 224 zone, and upward in elevation into the subalpine fir zone on the south side of Mt. Adams. 225 Increased drought will reduce forest productivity and favor more frequent and larger fires. 226 • Adaptation options: Incorporating an expectation of increased wildfire frequency and extent 227 into management plans can help managers better prepare for altered conditions. In dry forest 228 types, fuel management can help promote desired effects of fire. Thinning to reduce stand 229 densities, combined with prescribed fire, can alter fuel conditions, increase tree vigor, and 230 increase resilience to drought and insect outbreaks. It may be necessary to actively manage, 231 protect, and develop late-successional forest structure to maintain desired levels of late- 232 successional habitat on the landscape. Managers may need to reconsider genetic 233 characteristics of trees and include nursery stock from multiple seed zones when planting. 234 235 Special habitats— 236 • Effects: Ecosystem responses to climate change are expected to affect wildlife select 237 botanical species through altered habitat characteristics (food availability, nesting and resting 238 structures, water sources), competition, and predator-prey dynamics. Despite the flexibility 239 and adaptive capacity of many species, shifts in species ranges and local extirpation of some 240 species may result from climate change in combination with other stressors. Potential effects 241 of climate change on different focal habitats include: 242 Late-successional forest: Warmer temperature and reduced moisture availability in summer 243 will likely heighten competition among trees, increasing their susceptibility to insect attack. 244 An increase in high-severity fires and repeated fires in the same footprint could reduce the 245 extent of late-successional forest and physical features contributing to structural complexity. 246 The northern spotted owl, marbled murrelet, and fisher are a few species of concern that may 247 be sensitive to these potential changes. 248 Early-seral pre-forest: Although area of early-seral pre-forest is projected to increase, 249 reduced snowpack and summer precipitation may result in lower soil moisture, lower 250 establishment rates, reduced botanical species richness, and reduced site productivity in some 251 areas. DRAFTReduced species richness and lower site productivity, in turn, could reduce food Hudec et al. Front Matter 8

252 resources for some wildlife species. Elk and black-tailed deer are expected to be relatively 253 insensitive, whereas black bears are expected to be moderately sensitive. 254 Bigleaf maple: Increased disturbance associated with climate change may lead to increased 255 abundance of sprouting deciduous hardwoods, including bigleaf maple. Reduced snowpack, 256 longer dry seasons, and increased incidence of summer drought could affect habitat moisture 257 requirements for associated species (e.g., Puget Oregonian snail), but bigleaf maple is 258 expected to persist and potentially spread in some locations. 259 Quaking aspen: Aspen trees exhibit sensitivity to increasing temperatures, decreased 260 moisture availability, and altered fire regimes; so extent of aspen habitat, particularly in 261 upland areas, may decrease with climate change. However, aspen is a competitive post-fire 262 colonizer, and increased area burned by wildfire may offset threats to aspen. Aspen will 263 likely to continue to occur in areas expected to remain relatively moist, such as lacustrine 264 wetlands, meadows and wetlands fed by groundwater, and meadows in deep valleys. 265 Subalpine parkland: Climate change is predicted to adversely affect whitebark pine and some 266 subalpine parkland habitats. Clark’s nutcrackers are instrumental in whitebark pine 267 regeneration after large fires, and both species will be sensitive to a reduction in subalpine 268 parkland habitat. 269 Giant chinquapin: Primary threats to giant chinquapin are competition from overtopping 270 conifers, harvest, and conversion of forest land to other uses. Species that are drought 271 tolerant and fire resistant, including chinquapin, may become more competitive in a warmer 272 climate with more wildfire, particularly low- to mixed-severity fire. 273 Oregon white oak: Increased drought and disturbance may facilitate sprouting of deciduous 274 hardwoods including Oregon white oak. More frequent wildfires could also reduce 275 herbaceous biomass and favor Oregon white oak reproduction. 276 Meadows: Wet meadow habitat will likely decrease in a warmer climate because of projected 277 changes in hydrology, including more precipitation falling as rain, decreased snowpack, and 278 earlier spring snowmelt. Lower snowpack and a longer growing season in alpine and 279 subalpine wet meadows would encourage tree establishment on meadow perimeters. Loss of 280 high-elevation meadows would reduce habitat for wolverine, Cascade red fox, and American 281 pika. Climate change favors dry meadows, which are generally well-adapted to warm,dry 282 conditions and periodic drought. Increased fire frequency and drought limitations on tree 283 species distribution may increase area and quality of dry meadow habitat. Mardon skipper 284 will be sensitive to changes in dry, grassy meadows. 285 Rock outcrops: Higher summer temperatures and lower summer moisture could negatively 286 affect mountain goat populations on rock outcrops by reducing the amount and quality of 287 late-season forage. 288 Rocky balds: A warmer climate with more drought may have positive effects on rocky balds 289 by limiting conifer establishment and growth.. Increased fire frequency and extent could 290 expand the area of balds by killing small trees on the margins. 291 Alpine: In a warmer climate, alpine habitat may gradually migrate to higher elevations in 292 some locations. However, changes in climate may occur at a rate that exceeds migration 293 capacity for some species. American pika, hoary marmot, Cascade red fox, and wolverine 294 will be sensitive to changes in alpine habitat. 295 Riparian: These areas provide critical habitat to a diverse array of species, including 296 Neotropical birds, ducks, amphibians, and rare botanical species such as cold-water 297 corydalisDRAFT. Summer streamflows may decrease with warming climate because of earlier Hudec et al. Front Matter 9

298 snowmelt. Increasing temperatures and evapotranspiration and decreasing summer 299 streamflows may lead to drying in some riparian areas, particularly on the eastside of the 300 . Changes in riparian plant species composition and reduced riparian extent 301 could result in direct losses to the quantity and quality of ecological contributions of riparian 302 vegetation, such as wildlife habitat, shade over streams, and maintenance of water quality. 303 Wetlands and groundwater-dependent ecosystems: Wetlands are expected to be highly 304 vulnerable to climate change as a function of altered snowpack, precipitation regimes, and 305 groundwater recharge and discharge. Warming in all seasons and reduced summer 306 precipitation would result in increased evapotranspiration, decreased soil moisture in 307 summer, earlier drawdown, and reduced minimum water levels in wetlands. A warmer 308 climate may negatively impact Oregon spotted frog as seasonal drying of aquatic habitats and 309 altered vegetation become more common. 310 • Adaptation options: Assessing where late-successional forests are most at risk to fire and 311 insects will help prioritize actions such as fuel treatments and construction of fuel breaks. 312 Maintaining desired densities of native species, propagating drought-tolerant native species, 313 and controlling nonnative species may increase resilience in many habitat types.. 314 Decommissioning roads, reducing road connectivity, and redesigning drainage to increase 315 water retention would mitigate some habitat damage caused by management actions. 316 Maintaining or restoring stream channel form helps increase hydrologic function and store 317 water, which is beneficial for riparian and wetland vegetation, water quality, and aquatic 318 habitat. Increased monitoring will help land managers understand the ongoing effects of 319 climate change on special habitats. 320 321 Recreation— 322 • Effects: Demand for warm-weather activities such as hiking, camping, and nature 323 viewing is expected to increase because of the direct effect of a warmer climate on season 324 length. Earlier availability of snow-free sites and an increase in warm-weather days in 325 spring and autumn will increase access and challenge the ability of U.S. Forest Service 326 recreation staff to manage sites and activities. Extreme heat during summer months may 327 temporally shift demand to cooler weeks towards the shoulder season, or spatially shift 328 demand to more shaded sites. Climate change is expected to reduce opportunities for 329 snow-based winter activities, as the extent and duration of snowpack declines. Warming 330 temperatures may increase participation in hunting, birding, and viewing wildlife, 331 although extreme heat would have a negative effect. Forest products gathering patterns 332 may be altered due to warming temperatures and changing location and accessibility of 333 harvest sites. Water-based activities will likely be affected by lower and more variable 334 water levels in rivers, lakes, and reservoirs. 335 • Adaptation options: Capacity of recreation sites can be adjusted to meet increased 336 demand in shoulder and summer seasons (e.g., bigger campgrounds). Increased demand 337 for water-based recreation can be accommodated by managing lake and river access 338 capacity, and managing public expectations for site availability. Recreation management 339 will need to transition to shorter winter recreation seasons and changing use patterns. 340 Understanding the changing patterns of use will inform adjustments that can increase the 341 capacity of recreation sites. Managers may need to pay particular attention to road access 342 duringDRAFT shoulder seasons to prevent damage by vehicles. Recreation access near riparian Hudec et al. Front Matter 10

343 areas and lake shores may need to be managed more intensively to reduce human 344 impacts. 345 346 Ecosystem services— 347 • Effects: Higher temperature and increased frequency and extent of disturbances may alter 348 forest structure and growth, thus affecting both timber supply and carbon sequestration. 349 Biophysical changes may also have implications for local and global socioeconomic 350 conditions as well, affecting industries and communities that depend on harvest of timber and 351 other forest products. The ability of forests to sequester carbon will likely decrease if warmer 352 climate increases physiological stress in trees and increases the frequency and extent of 353 disturbances. The effects of increased wildfire are a big potential concern for air quality. 354 Climate change may also affect biophysical structures, processes, and functions related to 355 cultural resources, including “first foods” (e.g., huckleberries, salmon) valued by Native 356 Americans and others. 357 358 SWAP facilitated a climate change adaptation effort that achieved specific elements of 359 the U.S. Forest Service climate change strategy and provided an improved scientific context for 360 resource management, planning, and ecological restoration in Southwest Washington. The 361 adaptation options developed, many of which are already components of current management 362 practices, provide a pathway for slowing the rate of deleterious change in resource conditions. 363 Timely implementation of adaptation in resource planning and management will help maintain 364 critical structure and function of aquatic and terrestrial ecosystems in Southwest Washington. 365 Long-term monitoring will help detect potential climate change effects on natural resources of 366 concern and evaluate the effectiveness of adaptation options that have been implemented. 367 368 369 Acknowledgments 370 We thank leadership and resource managers of Gifford Pinchot National Forest for their support 371 of the Southwest Washington Adaptation Partnership. We thank all participants in the 372 vulnerability assessment and adaptation workshop for their enthusiasm and contributions. The 373 following reviewers provided insightful comments that greatly improved this publication: Nicole 374 DeCrappeo, John Stevenson, Joe Casola, Sam Brenkman, Correigh Greene, Karl Polivka, Miles 375 Hemstrom, Jane Kertis, Warren Devine, Kevin James, Shannon Claeson, Regina Rochefort, 376 Daniel Cressy, Megan Lawson, Nancy Brunswick, Steve Klein, Lesley Jantarasami, Delilah 377 Jaworski, Shiloh Halsey, Richard Bigley, and Brian Kittler. Amy Mathie provided outstanding 378 support with geospatial analysis and cartography. Funding was provided by the U.S. Forest 379 Service Pacific Northwest Research Station, Pacific Northwest Region, and Office of 380 Sustainability and Climate. Our hope is that the Southwest Washington Adaptation Partnership 381 will maintain an ongoing dialogue about climate change in the years ahead, catalyzing activities 382 that promoteDRAFT sustainability in the region. Hudec et al. Chapter 1 11

383 Chapter 1: Introduction 384 385 1 386 Joanne J. Ho 387 388 389 Introduction 390 391 The Southwest Washington Adaptation Partnership (SWAP) (fig. 1.1) is a science-management 392 partnership that includes Gifford Pinchot National Forest, U.S. Forest Service (USFS) Pacific 393 Northwest Research Station, USFS Pacific Northwest Region, University of Washington, 394 Washington Department of Natural Resources, and other local interest groups. Initiated in 2015, 395 the SWAP is a collaborative project with the goals of increasing climate change awareness, 396 assessing climate change vulnerability, and developing science-based adaptation options to 397 reduce adverse effects of climate change and ease the transition to new climate states and 398 conditions (see http://adaptationpartners.org/swap). Developed in response to proactive climate 399 change strategies of the USFS (USDA FS 2008, 2010a,c), and building on previous efforts in 400 national forests (Halofsky and Peterson 2017; Halofsky et al. 2011; 2018a, b; in press; Littell et 401 al. 2012; Raymond et al. 2013, 2014; Rice et al. 2012; Swanston et al. 2011, 2016), the 402 partnership brings together resource managers, research scientists, and stakeholders to plan for 403 climate change in Southwest Washington. 404 405 406 Gifford Pinchot National Forest 407 408 Gifford Pinchot National Forest was named after the first Chief of the Forest Service, who 409 actively supported environmental conservation as a philosophy and approach to forest 410 management. The Forest encompasses over 530,000 ha, including Mount St. Helens National 411 Volcanic Monument and approximately 73,000 ha of wilderness. Its location near the Portland- 412 Vancouver metropolitan area makes it a popular destination for visitors looking for a variety of 413 outdoor activities, including hiking and walking, bicycling, climbing, hunting, fishing, horseback 414 riding, camping, nature viewing, boating, and a number of winter sports such as skiing and 415 snowmobiling. Broad stretches of old-growth forest are home to the northern spotted owl (Strix 416 occidentalis), and a vast network of rivers and streams provide habitat for Chinook salmon 417 (Oncorhynchus tshawytscha), coho salmon (O. kisutch), steelhead (O. mykiss), and bull trout 418 (Salvelinus confluentus). Gifford Pinchot National Forest is one of 19 national forests affected by 419 the Northwest Forest Plan, which includes extensive standards and guidelines as part of a 420 comprehensive ecosystem management strategy. 421 422 423 Climate Change Response in the Forest Service 424

1 Joanne J. Ho is a research economist with the University of Washington, School of EnvironmentalDRAFT and Forest Sciences, Seattle, WA. Hudec et al. Chapter 1 12

425 Climate change is an agency-wide priority for the USFS, which has issued direction to 426 administrative units for responding to climate change (USDA FS 2008) (table 1.1). In 2010, the 427 USFS provided specific direction to the National Forest System in the form of the National 428 Roadmap for Responding to Climate Change (USDA FS 2010a) and the Performance Scorecard 429 (2011-2016) for Implementing the Forest Service Climate Change Strategy (USDA FS 2010a). 430 The overarching goal of the USFS climate change strategy is to “ensure our national forests and 431 private working lands are conserved, restored, and made more resilient to climate change, while 432 enhancing our water resources” (USDA FS 2010a). To achieve this goal, starting in 2011, each 433 national forest and grassland began using a 10-point scorecard system to report accomplishments 434 on ten elements in four dimensions: (1) increasing organizational capacity; (2) partnerships, 435 engagement, and education; (3) adaptation; and (4) mitigation and sustainable consumption. 436 Progress towards accomplishing elements of the scorecard was reported annually from 2011to 437 2016 by each national forest and grassland; all units were expected to accomplish 7 of 10 criteria 438 by 2015, with at least one “yes” in each dimension. 439 SWAP built on previous efforts in ecosystem-based management to address climate 440 change in the western and tiered efforts in Southwest Washington to that broader 441 context. Other efforts (table 1.2) have also demonstrated the success of science-management 442 partnerships to increase climate change awareness among resource managers and promote 443 climate change adaptation on federal lands. These previous assessments were intended to help 444 national forest managers identify where limited resources could be best invested to increase 445 watershed resilience to climate change. 446 The processes, products, and techniques used for several studies and other climate change 447 efforts on national forests have been compiled in a guidebook for developing adaptation options 448 for national forests (Peterson et al. 2011). The guidebook outlines four key steps to facilitate 449 adaptation in national forests: (1) become aware of basic climate change science and integrate 450 that understanding with knowledge of local conditions and issues (review), (2) evaluate 451 sensitivity of natural resources to climate change (rank), (3) develop and implement options for 452 adapting resources to climate change (resolve), and (4) monitor the effectiveness of on-the- 453 ground management (observe) and adjust as needed. SWAP is focused on implementation of the 454 principles and practices discussed in the guidebook. 455 456 457 Southwest Washington Adaptation Partnership Process 458 459 The SWAP geographic area includes Gifford Pinchot National Forest, Mount St. Helens National 460 Volcanic Monument, Washington Department of Natural Resources lands, and forested lands 461 managed by private timber companies. The SWAP process included: 462 • A vulnerability assessment of the effects of climate change on fisheries, forest vegetation, 463 special habitats, recreation, and ecosystem services. These resource sectors were selected 464 by local resource specialists based on current management concerns and challenges. 465 • Development of adaptation options that will help reduce negative effects of climate 466 change and assist the transition of biological systems and management to a warmer and a 467 changing climate. 468 • Development of an enduring science-management partnership to facilitate ongoing 469 dialogue and activities related to climate change. 470 VulnerabilityDRAFT assessments typically involve measures of exposure, sensitivity, and Hudec et al. Chapter 1 13

471 adaptive capacity (Parry et al. 2007), where exposure is the degree to which the system is 472 exposed to changes in climate, sensitivity is an inherent quality of the system that indicates the 473 degree to which it could be affected by climate change, and adaptive capacity is the ability of a 474 system to respond and adjust to the exogenous influence of climate. Vulnerability assessments 475 can be both qualitative and quantitative and focus on whole systems or individual species or 476 resources (Glick et al. 2011). Several tools and databases are available for systematically 477 assessing sensitivity of species (e.g., Case and Lawler 2016, Luce et al. 2014, Potter and Crane 478 2010). 479 Assessment teams used scientific literature and expert knowledge to assess exposure, 480 sensitivity and adaptive capacity and to identify key vulnerabilities for the identified resource 481 areas of concern. The process took place over 16 months and involved monthly phone meetings 482 for each resource-specific assessment team. Each assessment team identified key questions to 483 address, selected values to assess, and determined which climate change models and tools best 484 informed the assessment. In some cases, assessment teams conducted spatial analyses and ran 485 and interpreted models, selected criteria on which to evaluate model outputs, and developed 486 maps of model outputs and resource sensitivities. To the greatest extent possible, teams focused 487 on effects and projections specific to the region and used the finest scale projections that are 488 scientifically valid. 489 Focusing on Southwest Washington, scientists and resource managers worked 490 collaboratively to provide the scientific foundation for operationalizing climate change in forest 491 management planning and project implementation (Peterson et al. 2011; Raymond et al. 2013, 492 2014; Swanston et al. 2016). After identifying and assessing vulnerabilities for each resource 493 sector, scientists, land managers, and stakeholders convened at a workshop in April 2016 in 494 Vancouver, Washington to present and discuss findings of the vulnerability assessment and to 495 elicit ideas for adaptation options. Facilitated dialogue was used to identify key sensitivities and 496 adaptation options. Participants identified strategies (general approaches) and tactics (on-the- 497 ground actions) for adapting resources and management practices to climate change as well as 498 opportunities for implementing these adaptation actions into projects, management plans, 499 partnerships, and policies. Participants generally focused on adaptation options that could be 500 implemented given our current scientific understanding of climate change effects, but they also 501 identified research and monitoring that would benefit future efforts to assess vulnerability and 502 guide management practices. Facilitators captured information generated during the workshops 503 with worksheets adapted from Swanston et al. (2016). 504 This publication contains an overview on Southwest Washington biogeography, land-use 505 history, and climate, and one chapter for each of the resource sectors addressed in the 506 vulnerability assessment: fish and aquatic habitat, forest vegetation, special habitats, recreation, 507 and ecosystem services. An additional chapter summarizes adaptation strategies and tactics that 508 were compiled at the workshop. 509 Resource managers and other decisionmakers can use this publication in several ways. 510 First, the vulnerability assessment will provide information on climate change effects needed for 511 forest planning, environmental effects analyses, conservation strategies, and monitoring. Second, 512 climate change sensitivities and adaptation options developed at the broad scale provide the 513 scientific foundation for finer-scale assessments. We expect that over time, and as needs and 514 funding align, appropriate adaptation options will be incorporated into plans for specific 515 management units. Third, we anticipate that resource specialists will apply the information in 516 this assessmentDRAFT to forest management projects, thus operationalizing climate-informed resource Hudec et al. Chapter 1 14

517 management and planning. 518 Adaptation planning is an ongoing and iterative process. Implementation of adaptation 519 planning or actions may occur at any time, such as when managers revise USFS land 520 management plans and other planning documents, or after the occurrence of extreme events and 521 ecological disturbances (e.g., wildfire, flood). We focus on adaptation options for the USFS, but 522 information in this publication can be used by other land management agencies as well. Just as 523 the SWAP process has been adapted from previous vulnerability assessments and adaptation 524 planning efforts, other national forests and organizations can further adapt the SWAP process, 525 thus propagating climate-informed management across larger landscapes. 526 527 528 Towards an All-Lands Approach to Climate Change Adaptation 529 530 The USFS climate change strategy identifies the need to build partnerships and work across 531 jurisdictional boundaries when planning for adaptation. This concept of responding to the 532 challenge of climate change with an “all-lands” approach is frequently mentioned, but a process 533 for doing so is rarely defined. In addition to representatives from the USFS and Washington 534 Department of Natural Resources, several other agencies and organizations participated in the 535 SWAP workshop, including Washington Department of Fish and Wildlife, South Gifford 536 Pinchot Collaborative Group, Pinchot Partners, Cascade Forest Conservancy, Klickitat County, 537 Skamania County, Gifford Pinchot Accountability Group, Cowlitz Tribe, Yakama Nation, 538 private timber companies, and members of the public. This type of partnership enables a 539 coordinated and complementary approach to adaptation that crosses jurisdictional boundaries. 540 The SWAP also provides a venue for agencies to learn from the practices of others so that the 541 most effective adaptation options can be identified. 542 Risks and vulnerabilities associated with climate change, and gaps in scientific 543 knowledge and policy need to be assessed on a continual basis. Engaging employees, partners, 544 and the general public in productive discussions about climate change and adaptation is an 545 integral part of successfully responding to climate change. Furthermore, sharing climate change 546 information, vulnerability assessments, and adaptation strategies across administrative 547 boundaries will further enhance the success of climate change responses in Southwest 548 Washington. 549 550 551 Literature Cited 552 553 Case, M.J.; Lawler, J.J. 2016. Relative vulnerability to climate change of trees in western 554 North America. Climatic Change. 136: 367–379. 555 Furniss, M.J.; Roby, K.B.; Cenderelli, D. [et al.]. 2013. Assessing the vulnerability of 556 watersheds to climate change: results of national forest watershed vulnerability pilot 557 assessments. Gen. Tech. Rep. PNW-GTR-884. Portland, OR: U.S. Department of 558 Agriculture, Forest Service, Pacific Northwest Research Station. 559 Glick, P.; Stein, B.A.; Edelson, N.A., eds. 2011. Scanning the conservation horizon: a guide to 560 climateDRAFT change vulnerability assessment. Washington, DC: National Wildlife Federation. Hudec et al. Chapter 1 15

561 Halofsky, J.E.; Peterson, D.L., eds. 2017. Climate change vulnerability and adaptation in the 562 Blue Mountains. Gen. Tech. Rep. PNW-GTR-939. Portland, OR: U.S. Department of 563 Agriculture, Forest Service, Pacific Northwest Research Station. 564 Halofsky, J.E.; Peterson, D.L.; Dante-Wood, S.K. [et al.]. 2018a. Climate change 565 vulnerability and adaptation in the Northern Rocky Mountains. Gen. Tech. Rep. RMRS- 566 GTR-374. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 567 Mountain Research Station. 568 Halofsky, J.E.; Peterson, D.L.; Ho, J.J. [et al.]. 2018b. Climate change 569 vulnerability and adaptation in the Intermountain Region. Gen. Tech. Rep. RMRS-GTR-xxx. 570 Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 571 Station. 572 Halofsky, J.E.; Peterson, D.L.; Ho, J.J., eds. In press. Vulnerability of natural resources to 573 climate change in the South Central Oregon region. PNW-GTR-XXX. Portland, OR: U.S. 574 Department of Agriculture, Forest Service, Pacific Northwest Research Station. 575 Halofsky, J.E.; Peterson, D.L.; O’Halloran, K.A.; Hoffman, C.H. 2011. Adapting to climate 576 change at Olympic National Forest and Olympic National Park. Gen. Tech. Rep. PNW-GTR- 577 844. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 578 Research Station. 579 Hayward, G.D.; Colt, S.; McTeague, M.L. [et al.], eds. 2017. Climate change vulnerability 580 assessment for the Chugach National Forest and the Kenai Peninsula. Gen. Tech. Rep. PNW- 581 GTR-950. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 582 Research Station. 583 Janowiak, M.K.; Iverson, L.R.; Mladenoff, D.J. 2014. Forest ecosystem vulnerability 584 assessment and synthesis for northern Wisconsin and western upper Michigan: a report from 585 the Northwoods climate change response framework project. Gen. Tech. Rep. NRS-136. 586 Newton Square, PA: U.S. Department of Agriculture, Forest Service, Northern Research 587 Station. 588 Littell, J.S.; Peterson, D.L.; Millar, C.I.; O’Halloran, K.A. 2012. U.S. national forests adapt 589 to climate change through science-management partnerships. Climatic Change. 110: 269– 590 296. 591 Luce, C.; Lopez-Burgos, V.; Holden, Z. 2014. Sensitivity of snowpack storage to precipitation 592 and temperature using spatial and temporal analog models. Water Resources Research. 50: 593 9447–9462. 594 Morelli, T.L.; Yeh, S.; Smith, N.M. [et al.]. 2012. Climate project screening tool: an aid for 595 climate change adaptation. Res. Pap. PSW-RP-263. Albany, CA: U.S. Department of 596 Agriculture, Forest Service, Pacific Southwest Research Station. 597 Parry, M.L.; Canzianai, O.F.; Palutikof, J.P. [et al.], eds. 2007. Climate change 2007: 598 impacts, adaptation and vulnerability: a contribution of working group II to the fourth 599 assessment report of the Intergovernmental Panel on Climate Change. Cambridge, United 600 Kingdom: Cambridge University Press. 601 Peterson, D.L.; Millar, C.I.; Joyce, L.A. [et al.]. 2011. Responding to climate change in 602 national forests: a guidebook for developing adaptation options. Gen. Tech. Rep. PNW-GTR- 603 855. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 604 Research Station. 605 Potter, K.M.; Crane, B.S. 2010. Forest tree genetic risk assessment system: a tool for 606 conservationDRAFT decision-making in changing times. Version 1.2. Asheville, NC: U.S. Hudec et al. Chapter 1 16

607 Department of Agriculture, Forest Service, Eastern Forest Environmental Threat Assessment 608 Center. http://www.forestthreats.org/current-projects/project-summaries/genetic-risk- 609 assessment-system. (9 February 2017). 610 Raymond, C.L.; Peterson, D.L.; Rochefort, R.M. 2013. The North Cascadia Adaptation 611 Partnership: a science-management collaboration for responding to climate change. 612 Sustainability. 5: 136–159. 613 Raymond, C.L.; Peterson, D.L.; Rochefort, R.M., eds. 2014. Climate change vulnerability and 614 adaptation in the North Cascades region, Washington. Gen. Tech. Rep. PNW-GTR-892. 615 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 616 Station. 617 Rice, J.; Tredennick, A.; Joyce, L.A. 2012. Climate change on the Shoshone National Forest, 618 Wyoming: a synthesis of past climate, climate projections, and ecosystem implications. 619 Gen.Tech. Rep. RMRS-GTR-274. Fort Collins, CO; U.S. Department of Agriculture, Forest 620 Service, Rocky Mountain Research Station. 621 Swanston, C.W.; Janowiak, M.K.; Brandt, L.A.; [et al.], eds. 2016. Forest adaptation 622 resources: climate change tools and approaches for land managers, 2nd edition. Gen. Tech. 623 Rep. NRS-GTR-87-2. Newtown Square, PA: U.S. Department of Agriculture, Forest Service, 624 Northern Research Station. 161 p. 625 Swanston, C.; Janowiak, M.; Iverson, L. [et al.]. 2011. Ecosystem vulnerability assessment 626 and synthesis: a report from the climate change response framework project in northern 627 Wisconsin. Gen. Tech. Rep. NRS-82. Newtown Square, PA: U.S. Department of Agriculture, 628 Forest Service, Northern Research Station. 629 U.S. Department of Agriculture, Forest Service [USDA FS]. 2008. Forest Service strategic 630 framework for responding to climate change. Version 1.0. Washington, DC: U.S. Department 631 of Agriculture, Forest Service, Climate Change Advisor’s Office. 632 http://www.fs.fed.us/climatechange/documents/strategic-framework-climate-change-1-0.pdf. 633 (9 February 2017). 634 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010a. A performance scorecard 635 for implementing the Forest Service climate change strategy. Washington, DC: U.S. 636 Department of Agriculture, Forest Service, Climate Change Advisor’s Office. 637 http://www.fs.fed.us/climatechange/pdf/performance_scorecard_final.pdf. (9 February 2017). 638 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010b. National roadmap for 639 responding to climate change. Washington, DC: U.S. Department of Agriculture, Forest 640 Service, Climate Change Advisor’s Office. 641 http://www.fs.fed.us/climatechange/pdf/roadmap.pdf. (6 June 2015). 642 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010c. Strategic plan 2010-2015 643 FY. Washington, D.C.: U.S. Department of Agriculture, Forest Service. 644 http://www.ocfo.usda.gov/usdasp/sp2010/sp2010.pdf. (9 February 2017). 645 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012. 2012 Planning Rule. 646 Washington, DC: U.S. Department of Agriculture, Forest Service. Federal Register. 77(68). 647 115 p. http://www.fs.usda.gov/Internet/FSE_DOCUMENTS/stelprdb5362536.pdf. (9 648 FebruaryDRAFT 2017). Hudec et al. Chapter 2

649 Biogeography, Land-Use History, and Climate in Southwest 650 Washington 651 652 1 653 Summer Kemp-Jennings, Jessica E. Halofsky and John B. Kim 654 655 656 Biogeography and Historical Land Use in Southwest Washington 657 658 The Southwest Washington Adaptation Partnership (SWAP) area covers unique landscapes in 659 Southwest Washington with a diverse cultural and natural history (see fig. 1.1). Gifford Pinchot National 660 Forest comprises the majority of the SWAP area, including Mount St. Helens National Monument, 661 seven wilderness areas (USDA FS 1990), and a portion of the White Salmon River (a designated Wild 662 and Scenic River) (USDA FS 2015a). The alpine peaks of Mt. Adams (3,743 m) and Mount St. Helens 663 (2,550 m) stand out above the forested ridges and steep river valleys in the area. Glacial and volcanic 664 activity influenced the geologic landscape (Franklin and Dyrness 1973), and the resulting 665 geomorphology of the region influences the distribution of terrestrial and aquatic ecosystems. 666 The SWAP area is located along the westside of the Cascade Range in close proximity to other 667 public and private lands, with access via several major roadways. Mt. Rainier National Park shares a 668 border to the north. The eastern boundary runs north to south through the crest of Mt. Adams, and is 669 shared with the Yakama Reservation, a portion of the Okanogan-Wenatchee National Forest, and the 670 town of Trout Lake, Washington. The Gorge, Mt. Hood National Forest, and Interstate 671 Highway 84 are directly south of the Gifford Pinchot National Forest. Interstate 5 and several towns, 672 including Longview, Kalama, and Mount Vista, are located to the west of the national forest. 673 Residents of several Washington and Oregon counties, including Skamania, Lewis, Cowlitz, 674 Clark, and Klickitat, are tied to the SWAP area through recreational visitation, employment related to 675 natural resources, and subsistence practices like hunting and collecting. Residents of Portland, 676 Vancouver, the Puget Sound area, and the Yakama Reservation also influence the study area (USDA FS 677 1990). 678 Human influence has shaped the landscape and ecology of the study area for thousands of years 679 (USDA FS 2015a). Like most of the Pacific Northwest, Native American influence is rich within the 680 SWAP area. The Yakama, Cowlitz, Nisqually, Puyallup, Squaxin Island, Klickitat, Wishram, Cascade, 681 Umatilla, and Steilacoom tribes all contribute to the cultural history of the area. Early Native Americans 682 hunted and gathered in the area, and after glacial retreat, settled and actively managed the land. For 683 example, huckleberry (Vaccinium spp.) fields were regularly burned to increase production (fig. 2.1). 684 During the early 1900s, Southwest Washington was a prominent summer gathering place for tribes. 685 Tribes from as far as Montana and Idaho would come to trade, hunt, gather berries, weave baskets, fish, 686 and participate in games and festivities. The Yakama Treaty of 1855, the Medicine Creek Treaty of

1 Summer Kemp-Jennings is a planning specialist with the Beaverhead-Deerlodge National Forest, 420 Barrett St., Dillon, MT 59725-3572; Jessica E. Halofsky is a research ecologist, University of Washington, College of the Environment, School of Environmental and Forest Sciences, Box 352100, Seattle, WA 98195-2100; and John B. Kim is a biological scientist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, Forestry Sciences Laboratory, 3200 SW Jefferson Way, Corvallis, OR 97331. DRAFT17

Hudec et al. Chapter 2

687 1854, the 1932 Handshake Agreement, and the 1997 Memorandum of Understanding all serve to outline 688 exclusive rights for Native American use of Gifford Pinchot National Forest land and resources, and to 689 provide a framework for cooperation between the tribes and the national forest (USDA FS 2015a). 690 The abundance of American beavers (Castor canadensis Kuhl) and other fur-bearing animals 691 along the rivers and streams of the area brought fur trappers from the British Hudson Bay Company in 692 the early 1800s. Fort Vancouver was established in 1824 and marked the first permanent non-indigenous 693 settlement in the area. By the end of the 1800s, homesteaders were farming in river valleys, grazing 694 cattle and sheep in meadows, and mining and logging throughout the region (USDA FS 2015a). In 1897, 695 the area was incorporated into Mt. Rainier Forest Reserve. The area went through a series of 696 reorganizations and name changes, including becoming Columbia National Forest in 1908, before 697 becoming the Gifford Pinchot National Forest in 1949. Some of the roads, trails, and buildings still in 698 use today were constructed by the Civilian Conservation Corps from 1933 to 1942 as part of a federal 699 program originating during the Great Depression. Demand for timber during the early 20th century 700 prompted intensive harvesting, planting, and fire suppression efforts throughout the forest. Current 701 demands on the forest still include timber harvesting, in addition to wildlife, fisheries, recreation, and 702 wilderness use (USDA FS 2015a). 703 Recent and historical glacial and volcanic activity is prominent in the current landscape. The 704 andesite and basalt flows from volcanic activity are a significant feature of the Cascade Range in 705 Southwest Washington. Only small pockets of igneous intrusive, sedimentary, and metamorphic rocks, 706 which dominate the North Cascades, are found in the SWAP area. The area is marked with remnant 707 lakes from the forces of small alpine glaciers during the Pleistocene era. Present-day perennial glaciers 708 are located on Mt. Adams, Mount St. Helens, and in the Goat Rocks Wilderness. 709 On May 18, 1980, the nine-hour eruption of Mount St. Helens dramatically changed the 710 surrounding landscape (USDA FS 2015b). The eruption produced the largest landslide in recorded 711 history, flattened 9.4 million m3 (4 billion board-feet) of timber, caused 57 human fatalities, destroyed 712 12 million salmon fry fingerlings in hatcheries, and reduced the elevation of Mount St. Helens from 713 2,950 m to 2,550 m (fig. 2.2). Mount St. Helens National Volcanic Monument was created in 1982. The 714 44,515 ha area of Mount St. Helens National Volcanic Monument is left to naturally respond to 715 disturbance and the environment and is designated for research, education, and recreation (Brantley and 716 Myers 2000). 717 The SWAP area contains stretches of the headwaters of 10 large rivers and 10 small rivers, 718 including the Nisqually, Cowlitz, Cispus, Green, Toutle, Kalama, Lewis, East Fork Lewis, Wind, Little 719 White Salmon, and White Salmon Rivers (USDA FS 1990). Within the river valleys, steep drops and 720 irregular terrain are common and result in an abundance of waterfalls and popular features for 721 whitewater sports. In 2005, 32 km of the Upper White Salmon River within Gifford Pinchot National 722 Forest were designated as Wild and Scenic (Wild and Scenic Rivers Act 1968, NWSRS 2015). The 723 White Salmon River is distinguished by clear, steady flows from springs and seeps, geologic features, 724 resident native fish, and importance as a Native American spiritual site. Four additional rivers in the 725 SWAP area have been recommended for Wild and Scenic designation, and 13 are being studied for 726 future recommendation (USDA FS 2015a). 727 Forests and vegetation of the SWAP area are important ecologically, economically, and 728 culturally. Forests dominated by Douglas-fir (Pseudotsuga menziesii [Mirb.] Franco) in Gifford Pinchot 729 National Forest are some of the most productive timber forests in the National Forest System. These 730 forests contain several plant species that are important commercially, recreationally, and culturally, 731 including common beargrass (Xerophyllum tenax [Pursh] Nutt.), huckleberries, and mushrooms DRAFT18

Hudec et al. Chapter 2

732 (Hummel et al. 2012, USDA FS 1990). Sixteen plants on the sensitive species list of the U.S. Forest 733 Service Pacific Northwest Region are found in the SWAP area (USDA FS 1990). Several habitat types 734 identified as potentially sensitive to climate change in previous vulnerability assessments in the Pacific 735 Northwest exist in the SWAP area, including old-growth forests, alpine tundra, riparian and wetland 736 vegetation, and subalpine parkland (Halofsky et al. 2011, Raymond et al. 2014). 737 Composition of plant communities in the SWAP area is typical of westside Pacific Northwest 738 forest: vegetation is dependent on moist conditions and differs from low to high elevation based on 739 tolerance to extreme temperatures. Douglas-fir dominated forests below 1,000 m are commonly 740 associated with western redcedar (Thuja plicata Donn ex D. Don), and western hemlock (Tsuga 741 heterophylla [Raf.] Sarg.). Forests from 1,000 m to 1,700 m include Pacific silver fir (Abies amabilis 742 Douglas ex J. Forbes), grand fir (A. grandis [Douglas ex D. Don] Lindl.) and noble fir (A. procera 743 Rehder). Mountain hemlock (T. mertensiana [Bong.] Carriére) and Engelmann spruce (Picea engelmanii 744 Parry ex Engelm.) are also found around 1,000 m and continue up to treeline. Subalpine fir (Abies 745 lasiocarpa [Hook.] Nutt.), lodgepole pine (Pinus contorta var. latifolia Engelm. ex S. Watson), western 746 white pine (Pinus monticola Douglas ex D. Don), and ponderosa pine (Pinus ponderosa Douglas ex P. 747 Lawson & C. Lawson) are present, but less common (USDA FS 1990). 748 The SWAP area supports a diverse array of terrestrial and aquatic animal species. Gifford 749 Pinchot National Forest is home to the largest concentration of Roosevelt elk (Cervus canadensis 750 roosevelti Erxleben) in the United States. Mountain goats (Oreamnos americanus de Blainville) inhabit 751 steep, rocky areas of alpine tundra in the region. The forest contains over 150 species that use or require 752 old-growth habitat (USDA FS 1990). Some of the old-growth-dependent species like the northern 753 spotted owl (Strix occidentalis caurina Merriam) are expected to be especially sensitive to changes in 754 climate (Carroll 2010). Abundant cold streams provide habitat for resident and anadromous fish species. 755 Multiple dams outside of the SWAP area restrict anadromous species to the Green River, East Fork 756 , and . The Lewis River is home to record-size steelhead (Oncorhynchus mykiss 757 Walbaum) (USDA FS 1990). 758 Wildfires are important components of disturbance in the SWAP area. Fires over the past century 759 have burned vast amounts of forest land. Large burns in the early 20th century, and throughout the last 760 15 years, are significant in the fire history and ecology of the area. The Yacolt burn in 1902 was the 761 largest in Washington history until the recent fire era, consuming 90,000 ha (McClure 2005). During the 762 20 years following the Yacolt burn, the area experienced a series of other large burns, including the 763 24,000-ha Cispus fire and a series of significant reburns to the original Yacolt burn area (Mack 2002). 764 Managing the landscape for fire continues to be an important issue today. Over the past 15 years, 765 nearly 30,000 ha have burned within, and surrounding, the SWAP area. The largest fires include the 766 3,237-ha Cold Springs fire of 2008, the 8,132-ha Cascade Creek fire of 2012, and the 21,877-ha Cougar 767 Creek fire of 2015 (NWCC 2015). 768 Native and nonnative insects and pathogens typical of westside Cascade forest reside 769 within the forest. Areas of highest mortality are located on the south side of Mt. Adams and 770 mainly attributed to the fir engraver (Scolytus ventralis LeConte), pine bark beetles 771 (Dendroctonus ponderosae Hopkins, Dendroctonus brevicomis LeConte and Ips spp. De Geer), 772 western balsam bark beetle (Dryocoetes confusus Swaine), and Douglas-fir beetle (Dendroctonus 773 pseudotsugae Hopkins). Recent mortality from Douglas-fir beetles significantly increased across 774 the study area, probably from outbreaks generated by down woody debris from extensive winter 775 storm damage (Dizic et al. 2014). The spruce beetle (Dendroctonus rufipennis Kirby) has been a 776 significant agent of mortality in recent years, with associated Engelmann spruce (Picea DRAFT19

Hudec et al. Chapter 2

777 engelmannii Parry ex Engelm.) mortality. Historically, extensive insect and disease outbreaks 778 and mortality have generally occurred in dense forest stands (USDA FS 2004). 779 Five resource issues strongly influence decisionmaking and policy within the SWAP 780 area: (1) recreation (especially trails and semi-primitive recreation opportunities), (2) Wild and 781 Scenic Rivers, (3) old-growth forests, (4) wildlife (especially northern spotted owl and cavity 782 nesters), and (5) forest (and timber) management. Many other resource issues are also addressed 783 here, but most are linked to these five issues. Old-growth and riparian ecosystems are 784 particularly important in current forest planning because of their wildlife habitat value. Human 785 visitors are considered an integral part of ecosystems, so access for recreation, economic and 786 cultural activities are important aspects of forest planning (USDA FS 1990). 787 Timber production is a fundamental element of the area’s history and current economy. Gifford 788 Pinchot National Forest has historically been one of the highest producers of timber in the 789 National Forest System. High timber yields provide employment in logging and manufacturing, 790 and 25 percent of timber sale receipts are returned to the counties in lieu of taxes (USDA FS 791 1990). During peak harvest in the 1970s and 1980s, the forest averaged 700,000-940,000 m3 792 (300-400 million board-feet) of timber harvest annually. New regulations during the 1990s, 793 especially the Northwest Forest Plan (USDA and USDI 1994), resulted in significant reductions 794 in timber harvest (Hirt 1999). Currently, the forest annually harvests 70,000-94,000 m3 (30-40 795 million board-feet). The current amount of timber harvested is proactively regulated to minimize 796 negative impacts on wildlife habitat, scenery, recreation, transportation systems, and water 797 quality. Specific considerations include retention of snags and large trees for cavity excavators 798 and species that depend on old-growth and mature forest, and retention of thermal cover for deer, 799 elk, and mountain goats (USDA FS 1990). 800 The area provides abundant recreation opportunities for over 3 million people who reside 801 within a two-hour drive. Visitors are drawn by the wide range of available activities, from 802 primitive backpacking and mountaineering to car camping and sightseeing. Popular scenery 803 includes old-growth trees, snow-capped mountains, glaciers, lakes, streams, waterfalls, and rock 804 outcrops (USDA FS 1990). The most popular recreation activities are viewing natural features, 805 hiking and walking, hunting, and driving for pleasure (USDA FS 2009-2014). 806 807 808 Historical Climate in Southwest Washington 809 810 The climate of the SWAP area is mainly maritime and heavily influenced by the Pacific Ocean 811 and , with secondary orographic effects of the Cascade Range. Summers 812 are relatively cool and dry and winters are relatively mild and wet. Annual precipitation ranges 813 from 1,520 mm in the Cowlitz valley to 3,040 mm at higher elevations at the crest of the 814 Cascades (USDA FS 1990). Temperatures also vary considerably with elevation. 815 Several different analyses suggest that temperatures have increased over the last century 816 in Gifford Pinchot National Forest. An analysis based on PRISM gridded climate data (Daly et 817 al. 2001) suggests that temperatures have increased since 1895 (by 0.4 °C per century) and that 818 average annual temperatures have generally been above the 20th century average of 6.4 °C since 819 2000, with only three years below the 20th century average (fig. 2.3). In the western mountains, 820 PRISM has been shown to have an artificial amplification in warming trend (Oyler et al. 2015). 821 However, climate division data (NOAA National Centers for Environmental Information 2017), DRAFT20

Hudec et al. Chapter 2

822 specifically for the Cascade Mountains West Climate Division, also suggest that temperatures 823 have warmed in the region (0.8 °C per century). Temperatures for the larger Pacific Northwest 824 region have also increased (0.7 °C between 1985 and 2011; Mote et al. 2013). 825 There have been no significant annual precipitation trends in the Gifford Pinchot National 826 Forest or the surrounding area since 1895, based on PRISM gridded climate data (fig. 2.4) and 827 data for the Cascade Mountains West Climate Division (data not shown) (NOAA National 828 Centers for Environmental Information 2017). Regional analyses have indicated that spring 829 precipitation has increased, although the trend is not necessarily tied to increased greenhouse gas 830 concentrations (Abatzoglou et al. 2014). Precipitation in the Pacific Northwest is still dominated 831 by interannual variability, driven by phenomena such as the El Niño Southern Oscillation 832 (ENSO) (Abatzoglou et al. 2014, Mote et al. 2013). For example, in El Niño years, the Pacific 833 Northwest tends to be warmer and drier (Abatzoglou et al. 2014), sometimes leading to drought 834 and higher area burned by wildfire. However, climate datasets, including PRISM, rely on 835 weather stations, most of which are located at lower elevations. Luce et al. (2013) suggest that 836 reduced upper-level wind speeds since 1950 may have led to declines in mountain precipitation 837 in the region. 838 839 840 Projected Future Climate in Southwest Washington 841 842 Atmospheric scientists use global climate models (GCMs) to model Earth’s climate. Many 843 modeling groups have developed and run GCM simulations, which project future global climate 844 under different future scenarios. The Coupled Model Intercomparison Project (CMIP) is a 845 coordinated experiment involving many of these modeling groups worldwide, offering many 846 simulations for scientists to assess the range of future climate projections for the globe. The latest 847 CMIP experiment is the fifth phase of the project, referred to as CMIP5 (Taylor et al. 2009). 848 For CMIP5, simulations of future climate were driven by scenarios called Representative 849 Concentration Pathways (RCPs), which were created using a different process than those used to 850 create the previous generation of climate change scenarios, the Special Report on Emissions 851 Scenarios (SRES) (Nakićenović and Swart 2000). The SRES scenarios were designed by 852 beginning with future socio-economic scenarios, whereas RCPs were designed by defining future 853 radiative forcing values, and then developing socio-economic scenarios and natural system 854 dynamics that are consistent with the radiative forcing values. RCPs encompass the range of 855 current estimates regarding the evolution of radiative forcing, or the assumed rate of extra energy 856 entering the climate system throughout the 21st century and beyond (van Vuuren et al. 2011). 857 More information on CMIP can be found at https://www.wcrp-climate.org/wgcm-cmip. 858 GCMs simulate the climate at spatial scales too coarse for many types of regional studies. 859 Several climate research groups use advanced techniques to downscale spatially coarse climate 860 data to finer spatial scales. To explore potential future climate in the SWAP region, we used 861 NASA Earth Exchange Downscaled Climate Projections (NEX-DCP30), which comprises over 862 30 CMIP5 GCM projections downscaled to 30 arc seconds (approximately 800 m) spatial 863 resolution using the bias-correction spatial disaggregation method (Thrasher et al. 2013). We 864 focused on climate projections based on three GCMs: HadGEM2-ES, CSIRO-Mk3.6.0, and 865 NorESM1-M. Climate projections under RCP 8.5—among the most aggressive scenarios of 866 increasing greenhouse gases to the end of the 21st century—were summarized for a southwest DRAFT21

Hudec et al. Chapter 2

867 Washington area (fig. 2.5) for a historical period (1980-2009) and three future time periods: early 868 century (2010-2039), mid-century (2040-2069) and late century (2070-2099). Variables analyzed 869 included mean annual temperature, mean monthly temperature, mean annual precipitation, and 870 mean monthly precipitation for each of the four time periods. 871 For the southwest Washington area, all three GCMs show an increase in temperatures in 872 the future under RCP 8.5 (fig. 2.6). For early century, models project warming of 1.3-2.3 °C 873 compared to 1950-1979. Models project warming of 2.5-4.2 °C for mid-century and 4.3-6.4 °C 874 for late-century compared to 1950-1979. Seasonally, the largest increases in temperature are 875 projected for summer (+5.7-6.8 °C on average across models for June-August in 2070-2099; fig. 876 2.7). 877 Temperature projections for the southwest Washington area (between 2.5 °C and 4.2 °C 878 by 2070 under RCP 8.5) are consistent with warming projections for the Pacific Northwest 879 region of 1.1 °C to 4.7 °C (for 2041-2070 compared to 1950-1999), with the lower end possible 880 only if global greenhouse gas emissions are significantly reduced (RCP 4.5) (Mote et al. 2013). 881 As for season projections, temperature increases are expected to be greatest in summer across the 882 Pacific Northwest (Mote et al. 2013). In all analyses for the Pacific Northwest, there is no GCM 883 or scenario that suggests cooling in the future (Mote et al. 2013). 884 Precipitation projections are less certain than those for temperature, and projections for 885 future annual precipitation in the southwest Washington area range from wetter (NorESM1-M 886 model) to drier (HadGEM2-ES model) (fig. 2.8). Although projections vary among models, 887 mean summer precipitation is projected to decrease from 162 mm historically to 87-121 mm at 888 the end of the century, depending on the GCM (fig. 2.9). Projections for the Pacific Northwest 889 similarly show potential decreases in summer precipitation, with annual projections indicating 890 small trends compared to natural year-to-year variability (Mote et al. 2013). Model projections 891 for the region also agree that extreme precipitation events will likely increase in the future (Mote 892 et al. 2013). 893 894 895 Literature Cited 896 897 Abatzoglou, J.T.; Rupp, D.E.; Mote, P.W. 2014. Seasonal climate variability and change in the 898 Pacific Northwest of the United States. Journal of Climate. 27: 2125–2142. 899 Brantley, S.; Myers, B. 2000. Mount St. Helens – from the 1980 eruption to 2000. U.S. 900 Geologic Survey Fact Sheet 036-00. Online version 1.0. http://pubs.usgs.gov/fs/2000/fs036- 901 00/. (20 December 2015). 902 Carroll, C. 2010. Role of climatic niche models in focal-species-based conservation planning: 903 assessing potential effects of climate change on northern spotted owl in the Pacific 904 Northwest, USA. Biological Conservation. 143: 1432–1437. 905 Daly, C., G.H. Taylor, W. P. Gibson [et al.]. 2001. High-quality spatial climate data sets for 906 the United States and beyond. Transactions of the American Society of Agricultural 907 Engineers. 43: 1957–1962. 908 Dizic, A.; Hersey, C.; Kohler, G.; Omdal, D. [et al.]. 2014. Forest health highlights in 909 Washington—2014. U.S. Department of Agriculture, Forest Service and Washington 910 Department of Natural Resources.

DRAFT22

Hudec et al. Chapter 2

911 http://file.dnr.wa.gov/publications/rp_fh_2014_forest_health_highlights.pdf. (15 December 912 2015). 913 Franklin, J.F.; Dyrness, T.C. 1973. Natural vegetation of Oregon and Washington. Gen. Tech. 914 Rep. PNW-GTR-8. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 915 Northwest Forest and Range Experiment Station. 417 p. 916 Halofsky, J.E.; Peterson, D.L.; O’Halloran, K.; Hawkins Hoffman, C. 2011. Adapting to 917 climate change at Olympic National Forest and Olympic National Park. Gen. Tech. Rep. 918 PNW-GTR-844. Portland, OR: U.S. Department of Agriculture, Forest Service. Pacific 919 Northwest Research Station. 144 p. 920 Hirt, P.W. 1999. Creating wealth by consuming place: Timber management on the Gifford 921 Pinchot National Forest. In: White, R; Findlay, J.M., eds. Power and place in the North 922 American west. Seattle: University of Washington Press: 204–232. 923 Hummel, S.; Foltz-Jordan, S.; Polasky, S. 2012. Natural and cultural history of beargrass 924 (Xerophyllum tenax). Gen. Tech. Rep. PNW-GTR-864. Portland, OR: U.S. Department of 925 Agriculture, Forest Service. Pacific Northwest Research Station. 80 p. 926 Luce, C.H.; Abatzoglou, J.T.; Holden, Z.A. 2013. The missing mountain water: slower 927 westerlies decrease orographic enhancement in the Pacific Northwest USA. Science. 342: 928 1360–1364. 929 Mack, C. 2002. The “Columbia National Burn”: A history of fires and firefighting on the 930 Gifford Pinchot National Forest. 931 http://www.fs.usda.gov/Internet/FSE_DOCUMENTS/fsbdev3_004866.pdf. (10 January 932 2016). 933 McClure, R. 2005. Washington’s “awful conflagration”—The Yacolt Fire of 1902. Fire 934 Management Today. 65: 1. 935 Mote. P.; Abatzoglou, J.; Kunkel, K. 2013. Climate change and variability in the past and 936 future. In: Dalton, M.M.; Mote, P.W.; Snover, A.K., eds. Climate change in the Northwest: 937 implications for our landscapes, waters and communities. Washington, DC: Island Press: 25– 938 40. 939 Nakićenović, N.; Swart, R., eds. 2000. Special report on emissions scenarios. A special report 940 of working group III of the Intergovernmental Panel on Climate Change. Cambridge and 941 New York: Cambridge University Press. 599 p. 942 National Oceanic and Atmospheric Administration [NOAA] National Centers for 943 Environmental Information. 2017. Climate at a glance. 944 http://www.ncdc.noaa.gov/cag/time-series/us. (18 August 2017). 945 National Wild and Scenic Rivers System [NWSRS]. 2015. White Salmon River, Washington. 946 http://www.rivers.gov/rivers/white-salmon.php. (25 November 2015). 947 Northwest Interagency Coordination Center [NWCC]. 2015. Large fire information 948 summary. http://gacc.nifc.gov/nwcc/information/fire_info.aspx. (12 January 2016). 949 Oyler, J.W.; Dobrowski, S.Z.; Ballantyne, A.P. [et al.]. 2015. Artificial amplification of 950 warming trends across the mountains of the western United States. Geophysical Research 951 Letters. 42: 153–161. 952 Raymond, C.L.; Peterson, D.L.; Rochefort, R.M., eds. 2014. Climate change vulnerability and 953 adaptation in the North Cascades region, WA. Gen. Tech. Rep. PNW-GTR-892. Portland, 954 OR: U.S. Department of Agriculture, Forest Service. Pacific Northwest Research Station. 955 279 p. DRAFT23

Hudec et al. Chapter 2

956 Taylor, K.E.; Stouffer, R.J.; Meehl, G.A. 2009. A summary of the CMIP5 experiment design. 957 https://www.wcrp-climate.org/wgcm-cmip/wgcm-cmip5. (6 December 2017). 958 Thrasher, B.; Xiong, J.; Wang, W. [et al.]. 2013. Downscaled climate projections suitable for 959 resource management. Eos Transactions American Geophysical Union. 94: 321–323. 960 U.S. Department of Agriculture, Forest Service [USDA FS]. 1990. Land and resource 961 management plan: final environmental impact statement. Vancouver, WA: Gifford Pinchot 962 National Forest. 963 U.S. Department of Agriculture, Forest Service [USDA FS]. 2004. Forest insect and disease 964 conditions report. Portland, OR: Pacific Northwest Region. 965 http://www.fs.usda.gov/detail/r6/forest-grasslandhealth/insects- 966 diseases/?cid=fsbdev2_027291#ins-ind. (15 December 2015). 967 U.S. Department of Agriculture, Forest Service [USDA FS]. 2009—2014. National Visitor 968 Use Monitoring (NVUM) results. Gifford Pinchot National Forest. 969 http://apps.fs.fed.us/nfs/nrm/nvum/results/A06003.aspx/Round3. (29 December 2015). 970 U.S. Department of Agriculture, Forest Service [USDA FS]. 2015a. Gifford Pinchot National 971 Forest. http://www.fs.usda.gov/main/giffordpinchot/. (27 November 2015). 972 U.S. Department of Agriculture, Forest Service [USDA FS]. 2015b. Mount St. Helens 973 National Volcanic Monument. 974 http://www.fs.usda.gov/detail/mountsthelens/home/?cid=stelprdb5199437. (29 December 975 2015). 976 U.S. Department of Agriculture, Forest Service; U.S. Department of the Interior, Bureau of 977 Land Management [USDA and USDI]. 1994. Record of decision for amendments to Forest 978 Service and Bureau of Land Management planning documents within the range of the 979 northern spotted owl. 74 p. [plus attachment A: standards and guidelines]. 980 van Vuuren, D.P.; Edmonds, J.; Kainuma, M. [et al.]. 2011. The representative concentration 981 pathways: an overview. Climatic Change. 109: 5–31. 982 Wild and Scenic Rivers Act. 1968. 16 U.S.C. 1271 et seq. 983

DRAFT24

Hudec et al. Chapter 3

984 Chapter 3: Climate Change, Fish, and Aquatic Habitat in 985 Southwest Washington 986 987 Daniel Isaak, Ruth Tracy, Dona Horan, Jessica Hudec1 988 989 990 Introduction 991 992 993 Lands administered by Gifford Pinchot National Forest (GPNF) provide important 994 aquatic habitats for native cold-water fish species that have declined over the last 150 years as a 995 result of habitat fragmentation and degradation, overharvest, invasive species, interactions with 996 hatchery fish, and water development projects (Hessburg and Agee 2003, Nehlsen et al. 1991, 997 Sanderson et al. 2009). In recent decades, human-caused climate change has emerged as an 998 additional potential stressor to these fish populations. Warming air temperatures and changing 999 precipitation patterns have resulted in warmer stream temperatures (Bartholow 2005; Isaak et al. 1000 2017a, 2018; Petersen and Kitchell 2001), alterations to stream hydrology (Hamlet and 1001 Lettenmaier 2007, Hamlet et al. 2007, Luce et al. 2013), and changes in the frequency, 1002 magnitude, and extent of extreme events such as floods, droughts, and wildfires (Luce and 1003 Holden 2009, Marlier et al. 2017, Rieman and Isaak 2010). 1004 Biological evidence exists of fish population responses to those trends in the form of 1005 shifting spatial distributions (Al-Chokhachy et al. 2016, Eby et al. 2014), phenological 1006 adjustments (Crozier et al. 2011, Martins et al. 2012), and evolutionary change (Kovach et al. 1007 2012, Manhard et al. 2017). Notably, cold-water salmon and trout populations that are often of 1008 management and conservation concern show evidence of heat-related stress in some rivers 1009 during warm summers that may lead to fishing season closures, migration delays, and mortality 1010 events (Bowerman et al. 2016, Cooke et al. 2004, Keefer et al. 2009, Lynch and Risley 2003) as 1011 the fish aggregate into or seek cold micro-refugia (Ebersole et al. 2001, Torgersen et al. 1999). 1012 Continuation, and possible acceleration, of these trends during the 21st century (chapters 1 and 2) 1013 is likely to have important implications for the distribution, abundance, and persistence of some 1014 populations of fish species and will complicate conservation and management efforts on their 1015 behalf. 1016 Adapting to the challenges that climate change poses for cold-water fishes ultimately 1017 requires detailed information about local climatic conditions, trends, and target species status to 1018 assist in strategic and tactical decisionmaking. Therefore, rather than reviewing the large and 1019 growing literature that describes the many possible interactions among climate change and 1020 aquatic environments (Comte et al. 2013, Hauer et al. 1997, Hotaling et al. 2017, Isaak et al.

1 Daniel J. Isaak is a research fish biologist, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, 322 East Front Street, Suite 401, Boise, ID 83702; Ruth Tracy is a soil-water program manager, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, 1501 E. Evergreen Blvd. Vancouver, WA 98661; Dona Horan is a research fish biologist, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, 322 East Front Street, Suite 401, Boise, ID 83702; Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams RangerDRAFT District, 2455 Hwy 141, Trout Lake, WA 98650. 25

Hudec et al. Chapter 3

1021 2012, ISAB 2007, Lynch et al. 2016, Mantua et al. 2010, Mote et al. 2003, Rieman and Isaak 1022 2010, Whitney et al. 2016), we instead summarize and present information specific to the 1023 Southwest Washington Adaptation Partnership (SWAP) assessment. 1024 First, we provide a historical perspective of the aquatic habitats and past activities that 1025 affect their current status and ability to support aquatic species. Second, we describe the spatial 1026 extent of the stream and river habitats in the analysis area using high-resolution scenarios and 1027 then describe climate related historical trends and future projections in hydrologic and thermal 1028 regimes. Third, we describe the status and potential climate vulnerabilities for fish species of 1029 concern in the analysis area, which were identified from discussions with land managers and 1030 U.S. Forest Service (USFS) regional staff at the outset of the SWAP assessment. Species were 1031 chosen based on their perceived vulnerability to climate change and because of their societal 1032 prominence as species listed as endangered or threatened under the U.S Endangered Species Act 1033 (ESA), including Chinook salmon (Oncorhynchus tshawytscha Walbaum in Artedi; spring and 1034 fall runs), coho salmon (O. kisutch Walbaum), the anadromous form of rainbow trout, commonly 1035 referred to as steelhead (O. mykiss Walbaum; summer and winter runs), and bull trout (Salvelinus 1036 confluentus Suckley) (table 3.1). Fourth, this information is discussed within the context of 1037 potential climate adaptation options that could alleviate future stresses for the species of concern. 1038 1039 1040 Aquatic Landscape Conditions 1041 1042 The aquatic assessment area for Southwest Washington contains numerous streams, rivers, and 1043 lakes that drain topographically complex, forested lands of mixed private and federal ownership. 1044 Portions of 12 major river systems occur within GPNF and include headwaters for the Cispus, 1045 Cowlitz, East Fork Lewis, Green, Kalama, Lewis, Little White Salmon, Muddy, Nisqually, 1046 Toutle, White Salmon, and Wind rivers. Many geological barriers and a few dams limit 1047 anadromous fish movements across the landscape (fig. 3.1), so only streams in the Green River, 1048 , and Wind River drainages are currently occupied. Dams outside the 1049 GPNF are barriers to anadromous runs on portions of other rivers that include the Nisqually, 1050 Cowlitz, Cispus, and Lewis Rivers. Dam removals at Hemlock Lake and Martha Creek have 1051 recently increased access to historical anadromous fish habitats in the Wind River drainage and 1052 the Condit Dam removal increased access to portions of the White Salmon River downstream of 1053 the GPNF boundary. 1054 Rivers and streams in southwest Washington once hosted highly productive salmon and 1055 steelhead fish populations and fisheries. Cool stream temperatures, clean gravel beds, and deep 1056 pools supported healthy aquatic systems and high levels of biological diversity (LCFRB 2010). 1057 Wild runs numbering a million or more fish were estimated to once occur in Lower Columbia 1058 River (LCR) streams of Washington, but runs now average about 30,000 per year (LCFRB 1059 2010). Details regarding the status and trends of populations within each of the four species are 1060 provided in numerous agency and recovery planning reports (Ford et al. 2011, LCFRB 2010, 1061 USFWS 2015). 1062 Alteration of aquatic habitats by human activities beginning in the late 19th century 1063 extensively degraded, fragmented, and simplified stream channels and floodplains within these 1064 systems. Growth of the timber industry was accompanied by development of an extensive road 1065 network that contributed fine sediments into streams, increased the incidence of hillslope 1066 failures, andDRAFT sometimes restricted fish movements where road culverts provided inadequate 26

Hudec et al. Chapter 3

1067 stream passage (Steel et al. 2004, Trombulak and Frissell 2000). Extensive road networks that 1068 accompanied harvest in some basins have also been shown to alter hydrologic regimes and 1069 increase peak flows compared to unharvested basins (Jones and Grant 1996, Moore and 1070 Wondzell 2005). 1071 Moving timber downstream to sawmills was often accomplished by temporary 1072 development and destruction of splash dams; log passage along stream courses was expedited by 1073 removal of large woody debris and other roughness elements that contributed to habitat diversity 1074 (Miller 2010, Sedell and Froggatt 1984, Wing and Skaugset 2002). Road construction and timber 1075 harvest adjacent to streams opened riparian canopies and probably contributed to alteration of 1076 stream thermal regimes (Holtby 1988, Johnson and Jones 2000, Moore et al. 2005). Repeat 1077 surveys spanning 1937-1987 also showed that channels in managed watersheds are often 1078 significantly wider than those in protected watersheds (Dose and Roper 1994) because of 1079 increased sediment loads, altered hydrology, and poor stream bank conditions, which may 1080 further exacerbate temperature increases (Beschta 1978). The GPNF currently has 142 km of 1081 temperature-impaired segments on 30 streams, per the Washington State Department of Ecology 1082 standards assessment. 1083 The depressed status of anadromous fish species in the assessment area has motivated 1084 prominent regional conservation efforts (Reeves et al. 2018, USDA FS 2005), and subsequent 1085 enactment of the Aquatic and Riparian Effectiveness Monitoring Plan (AREMP) to monitor 1086 stream conditions throughout the region (Reeves et al. 2003). Trend monitoring datasets 1087 collected since 1994 at the inception of the AREMP program suggest that stream conditions on 1088 the GPNF and the region’s other national forests have generally been stable or improving, 1089 changes which may be attributable to better management practices, reductions in timber harvest, 1090 and decommissioning of some roads (Lanigan et al. 2012). 1091 However, significant stream recovery is expected to take decades, given the extent of 1092 historical modifications, and habitats are likely to remain much less diverse and productive than 1093 pre-settlement conditions for the foreseeable future. Specific to GPNF, aquatic restoration efforts 1094 have been ongoing for over 20 years. Current aquatic restoration is spread among multiple 1095 watersheds, which all include anadromous habitats. The goal of aquatic restoration in these 1096 watersheds is to actively improve fish habitat. Other watersheds receive less active restoration, 1097 but all land-use activities are regulated by broad-level protections. Detailed effects assessments 1098 of these disturbances on habitats are included in Watershed Analysis (Furniss et al. 2010, 2013), 1099 the Watershed Condition Framework (USDA FS 2011), and recovery plans for individual ESA- 1100 listed fish. 1101 1102 1103 Stream Climate Trends 1104 1105 To describe stream climate trends and the extent of habitat available to species of concern, we 1106 delineated the SWAP assessment network using the 1:100,000-scale National Hydrography 1107 Dataset (NHD)-Plus Version 2, which was downloaded from the Horizons Systems website 1108 (http://www.horizon-systems.com/NHDPlus/index.php; McKay et al. 2012) and filtered by 1109 minimum flow and maximum stream slope criteria. Summer flow values predicted by the 1110 Variable Infiltration Capacity hydrologic model (VIC; Wenger et al. 2010) were obtained from 1111 the WesternDRAFT U.S. Flow Metrics website 27

Hudec et al. Chapter 3

1112 (http://www.fs.fed.us/rm/boise/AWAE/projects/modeled_stream_flow_metrics.shtml) and linked 1113 to NHD-Plus stream reaches. 1114 The network was filtered to exclude reaches with slope greater than 15 percent or those 1115 with minimum summer flows less than 0.006 m3s-1, which approximates a low-flow wetted 1116 width of 1 m (based on an empirical relationship developed in Peterson et al. [2013]), because 1117 fish occurrences are rare in these areas (Isaak et al. 2017b). The steepest headwater reaches are 1118 also prone to frequent large disturbances (e.g., post-wildfire debris torrents) that may cause local 1119 extirpations of fish populations (May and Gresswell 2004, Miller et al. 2003). Application of 1120 these criteria created the final 4,500 km network that served as the basis for subsequent analyses 1121 and summaries. Fifty-three percent of the network flowed through USFS lands whereas the 1122 remainder flowed through private and state lands (fig. 3.1). 1123 Scenarios representing mean August stream temperature were downloaded from the 1124 NorWeST website (www.fs.fed.us/rm/boise/AWAE/projects/NorWeST.html; Isaak et al. 2016a) 1125 and linked to reaches in the analysis network. NorWeST scenarios have a 1-km resolution and 1126 were developed by applying spatial stream network models to temperature records for 1,258 1127 summers of measurement at 344 unique stream sites that were collected by resource agencies 1128 within the assessment area (Isaak et al. 2017a). The predictive accuracy of the NorWeST model 1129 (cross-validated r2=0.91; cross-validated root mean square prediction error=1.0 °C), combined 1130 with substantial empirical support, provided a consistent and spatially balanced rendering of 1131 temperature patterns and thermal habitat for all streams. To depict temperatures during a baseline 1132 period, we used the S1 scenario that represented average historical conditions for 1993-2011. 1133 Mean August stream temperatures during this period were 11.5 °C, ranged from 3.8 to 27.2 °C 1134 throughout the network, and were usually cooler in streams flowing through national forest lands 1135 at higher elevations (table 3.2). 1136 Future stream temperature scenarios were also downloaded from the NorWeST website 1137 and chosen for the same climate periods (2030-2059, hereafter 2040s; 2070-2099, hereafter 1138 2080s) and emission scenario (A1B) as those used for the streamflow analysis in the hydrologic 1139 assessment (chapter 3). The future NorWeST scenarios used were S30 (2040s) and S32 (2080s), 1140 which account for differential sensitivity and slower warming rates of the coldest streams that are 1141 usually buffered by groundwater (Isaak et al. 2016b, Luce et al. 2014). Future August stream 1142 temperature increases relative to the baseline period of 2000 were projected to average 1.3 °C by 1143 the 2040s and 2.2 °C by the 2080s, which implies a warming rate of approximately 0.2-0.3 °C 1144 per decade (table 3.2, fig. 3.2). That rate is larger than the historical summer warming rate of 1145 0.11-0.23 °C per decade for the 40-year period of 1976-2015 that has been estimated from long- 1146 term monitoring records in the assessment area (fig. 3.3; Isaak et al. 2012, 2018). 1147 Here we briefly summarize a subset of ecologically relevant hydrologic metrics. These 1148 include mean and summer flow volumes that dictate habitat volume, as well as the frequency and 1149 magnitude of peak flow events during the winter and runoff periods that could scour redds or 1150 reconfigure channel habitats. Hydrologic conditions were represented using metrics predicted by 1151 the VIC model for the same climate scenarios and periods as those used for the NorWeST 1152 scenarios to provide consistent comparisons. 1153 Across the assessment area, the frequency of days with high winter flows is expected to 1154 increase from 11 days during the historical period to 13-14 days in the future, with slightly larger 1155 increases on GPNF lands at higher elevations (table 3.3, fig. 3.4). Increases in peak flows mirror 1156 that pattern, with slightly larger increases in reaches flowing through GPNF lands (increases of 1157 13.0-22.8DRAFT percent) than the assessment area as a whole (increases of 9.6-17.3 percent) (fig. 3.5). 28

Hudec et al. Chapter 3

1158 Mean annual flows are predicted to increase slightly (3.1-3.9 percent) while summer flows are 1159 predicted to decrease substantially by 40-65 percent (fig. 3.6) in response to decreases in future 1160 snowpack size and earlier runoff. Because future hydrological changes depend to an extent on 1161 context within the network, different fish species will experience differences from these averages 1162 throughout their habitats that encompass subsets of the Southwest Washington area. These 1163 changes are summarized for each of the four focal species in tables 3.4-3.7 and discussed in the 1164 subsequent section. 1165 In addition to long-term climate trends, climatic cycles in ocean productivity are 1166 recognized for their strong effects on growth and survival of anadromous fishes in the Pacific 1167 Northwest (Hare et al. 1999, Mantua et al. 1997). More specifically, ocean productivity varies 1168 through time in response to sea surface temperatures and the strength of coastal upwelling tied to 1169 regional climate cycles like the El Niño Southern Oscillation (ENSO), Pacific Decadal 1170 Oscillation (PDO), and more recently, the North Pacific Gyre Oscillation (NPGO; Kilduff et al. 1171 2015). Although these cycles most strongly affect anadromous fishes during their oceanic life 1172 stages, inland effects on temperature, precipitation, and hydrologic regimes that alter the quality 1173 and quantity of freshwater habitat are also evident (Kiffney et al. 2002, Mote et al. 2003). 1174 Recent research has documented a link between climate change and increasing variance 1175 in the north Pacific Oscillation (NPO; Di Lorenzo and Mantua 2016), which is the primary driver 1176 of the NPGO and PDO that explains most of the variation in Chinook salmon and coho 1177 recruitment along the west coast of North America (Kilduff et al. 2015, Mantua 2015). 1178 Consistent with this view, recent winters have shown NPO activity at record highs and the 1179 warmest sea surface temperature anomalies ever recorded in the northeast Pacific (i.e., “the 1180 blob”), suggesting that extremes in physical conditions linked to salmon survival rates may 1181 become more frequent in future decades (Bond et al. 2014, Di Lorenzo and Mantua 2016). 1182 1183 1184 Focal Species Status and Vulnerability 1185 1186 Climate-related trends in aquatic habitat interact with the status, ecology, habitat preferences, 1187 and climatic sensitivity of individual species and populations to determine their vulnerability. In 1188 this section, those vulnerabilities are discussed and contextualized with regards to the subsets of 1189 streams that constitute habitat for the four focal species as delineated on critical habitat maps 1190 developed by the U.S. Fish and Wildlife Service (USFWS) for bull trout and by the National 1191 Oceanic and Atmospheric Administration (NOAA) for coho salmon, Chinook salmon, steelhead 1192 trout. Geospatial data describing the critical habitats were downloaded from the agency websites 1193 (USFWS: https://www.fws.gov/pacific/bulltrout/; NOAA: 1194 http://www.westcoast.fisheries.noaa.gov/maps_data/endangered_species_act_critical_habitat.ht 1195 ml). 1196 1197 1198 Coho Salmon 1199 1200 Coho salmon once returned to spawn more broadly in streams of the assessment area, but dams 1201 have precluded access to upstream areas of the Cowlitz and Lewis Rivers, which historically 1202 supported very large, diverse, and productive runs of early and late coho (LCFRB 2010). As a 1203 result, cohoDRAFT salmon currently access and use 584 km of streams and rivers distributed throughout 29

Hudec et al. Chapter 3

1204 the analysis area (Ford et al. 2011) (table 3.4, fig. 3.7). Ocean productivity cycles affect growth 1205 and survival of coho salmon and the number of adults that annually return to spawn, as is the 1206 case for all the anadromous species considered here (Beamish and Mahnken 2001, Hare et al. 1207 1999). Coho adults leave the ocean after 1-3 years and migrate upstream from October through 1208 January, with variation in timing occurring among populations and individuals within 1209 populations. Migration distances to spawning areas are moderate in length, can be completed in a 1210 few weeks, and coho salmon usually spawn within one or two weeks of reaching the spawning 1211 grounds (Willis 1954). 1212 Spawning streams consist of small, unconfined, low-gradient tributaries to larger rivers 1213 (Burnett et al. 2007), and females deposit eggs in redds that are excavated from the substrate 1214 before dying. The eggs hatch after 6-7 weeks from late winter to early spring, and alevins remain 1215 in the substrate for another 6-7 weeks while the yolk sac is absorbed. After emerging from redd 1216 substrates, young coho salmon spend 1-2 years growing in their natal streams and exhibit a 1217 general preference for pools, alcoves, and beaver ponds rather than habitats with higher flow 1218 velocities like glides and riffles (Nickelson et al. 1992, Steel et al. 2016). Once juvenile fish 1219 reach lengths of 100-150 mm, they transform into smolts and migrate to the ocean from late 1220 March through July. 1221 The sensitivity of coho salmon to climate change depends on the portion of the life cycle 1222 considered. Low sensitivities are expected during the freshwater migrations of adults and smolts 1223 because these movements occur during months with relatively cool temperatures and high flows. 1224 However, resident juvenile life stages are likely to be adversely affected by continuation of long- 1225 term summer flow declines and temperature increases. Declines in mean summer flows of 40-65 1226 percent, if realized later this century, would equate to losing a similar amount of habitat and 1227 reduce potential population sizes by intensifying competition for food and space. As mean 1228 summer flows decrease, the probability of extreme low-flow years and drought increases (Luce 1229 and Holden 2009), as was the case in 2015 when record low flows and warm temperatures 1230 occurred along much of the Oregon and California coasts (Marlier et al. 2017), prompting broad 1231 societal concern about fish mortality and unprecedented closures of freshwater fishing seasons 1232 throughout the region (ODFW 2015). Long-term warming trends during the summer may create 1233 chronic stresses for juvenile coho salmon in stream reaches that are already near the species’ 1234 maximum thermal tolerances and could force gradual upstream range contractions. Temperature 1235 increases, by accelerating growth or egg incubation rates, also have the potential to 1236 desynchronize the developmental phenology of juveniles from the temporal availability of 1237 habitats (Holtby 1988). 1238 Increased channel disturbance may negatively affect coho salmon populations during 1239 incubation and rearing life stages. If climate change enhanced variability of ocean cycles (Bond 1240 et al. 2014, Di Lorenzo and Mantua 2016) results in higher or more intense precipitation, larger 1241 peak flows could scour redds or cause mortalities of newly emerged and weakly swimming 1242 alevins. Locations where scour could occur, however, are strongly context dependent at local and 1243 network scales (Goode et al. 2013, McKean and Tonina 2013, Shellberg et al. 2010), with 1244 steeper channels in confined valleys where structural habitat complexity is low showing higher 1245 probabilities of disturbance (Sloat et al. 2017). If wildfires become more common, juvenile life 1246 stages could also be negatively affected in the short-term by fine-sediment deposition and debris 1247 flows into the channel network. Over the longer term, however, those events could have 1248 beneficial effects by adding spawning gravels and large woody debris that may increase habitat 1249 diversity (BissonDRAFT et al. 2003, Dunham et al. 2003). 30

Hudec et al. Chapter 3

1250 Although coho salmon populations may not be acutely vulnerable to the effects of 1251 climate change at any one life stage, the pervasive nature of climate means that cumulative 1252 effects accrue over the course of the full life cycle and may lead to negative synergies (Crozier et 1253 al. 2008, Honea et al. 2016). For example, exacerbation of multiyear or decadal cycles of poor 1254 ocean conditions could depress numbers of returning adults, which then reproduce poorly in 1255 freshwater habitats subject to extreme drought, warm temperatures, and channel disturbances. 1256 Coho salmon populations, such as those of most anadromous fishes, are buffered by density- 1257 dependent responses, diverse life histories, and multiple age classes (i.e., the portfolio effect; 1258 Schindler et al. 2010) but more extreme environmental conditions, if synchronized over larger 1259 spatial scales and longer time periods, will pose novel challenges to species resilience. 1260 1261 1262 Chinook Salmon 1263 1264 Chinook salmon within the assessment area occupy many of the same river systems as coho 1265 salmon but have a more restricted distribution of approximately 359 kilometers (table 3.5, fig. 1266 3.8). These salmon consist of two variants, a spring run of fish that migrates upriver from May 1267 through July, and a fall run of fish that migrates later in the year from September through 1268 December (Ford et al. 2011). Both variants are large bodied (10-20 kg) and use habitats 1269 associated with larger streams and rivers than those used by coho salmon. Although spring run 1270 fish migrate earlier in the year than fall run fish, they typically use spawning areas further 1271 upstream and often hold in deep pools near spawning sites for extended periods prior to initiating 1272 redd construction in August and September (Ratner et al. 1997). 1273 Fall Chinook salmon spawn lower in most rivers and shortly after reaching the spawning 1274 grounds (Healey 1991). Eggs incubate overwinter, juvenile fish rear for several months (fall 1275 Chinook salmon) or years (spring Chinook salmon), and then most smolt and emigrate to the 1276 ocean during high flows in the spring or early summer. Outmigrating smolts are often preyed 1277 upon by populations of nonnative predators in the Columbia River or warmer portions of natal 1278 rivers, which probably becomes a larger source of mortality during later parts of each year’s 1279 migration as river temperatures warm and predator species become more active (Rieman et al. 1280 1991). Once in the ocean, Chinook salmon range widely and grow for 1-4 years before returning 1281 to their natal rivers to spawn (Healey 1991). 1282 The potential vulnerabilities of Chinook salmon to climate change are similar in some 1283 regards to coho salmon. Altered ocean conditions will exert broad effects on growth, survival, 1284 and numbers of returning adults (Beamish and Mahnken 2001, Hare et al. 1999). However, 1285 spring run Chinook salmon adults migrate upriver during warm summer months and often 1286 experience thermally stressful conditions, which may alter migration timing or temporarily stop 1287 migrations during peak temperatures when fish are forced to seek cold microrefugia (Keefer et 1288 al. 2009, Torgersen et al. 1999). Because spring run fish stage for long periods prior to spawning, 1289 thermal stress may accumulate and could adversely affect the viability of eggs or increase pre- 1290 spawn mortality rates in adults (Bowerman et al. 2016). 1291 Risks to Chinook salmon redds and incubating eggs from channel scour may be relatively 1292 low because the fish usually spawn in larger rivers where valleys are less confined and peak flow 1293 energy is dissipated across floodplains (McKean and Tonina 2013, Sloat et al. 2017). 1294 Vulnerability of fall Chinook juveniles is also expected to be low because they spend little time 1295 in freshwaterDRAFT prior to ocean outmigration. However, spring Chinook juveniles rear for a year or 31

Hudec et al. Chapter 3

1296 more and may be adversely affected by similar mechanisms as those affecting juvenile coho 1297 salmon, namely long-term declines in summer flows, temperature increases, and enhanced 1298 frequency and intensity of extreme events that will intensify competition for food and space or 1299 possibly result in direct mortality. 1300 1301 1302 Steelhead Trout 1303 1304 Steelhead is the anadromous form of rainbow trout and populations within the assessment area, 1305 occupying the largest extent of the anadromous species at approximately 900 km of streams 1306 (table 3.6, fig. 3.9). Populations consist of two variants: summer-run steelhead that migrate into 1307 freshwaters from May to October, and winter-run fish that migrate from November to March and 1308 are more extensively distributed. Spawning occurs from January through March (Quinn 2005), 1309 so early migrating summer steelhead adults reside in deep pools for extended periods while 1310 waiting to spawn (Baigun 2003). Females usually excavate redds in steeper streams with more 1311 confined valleys than those used by salmon (Burnett et al. 2007, Reeves et al. 1998). After 1312 hatching, the juveniles rear for one to three years near the natal areas before smolting and 1313 migrating to the ocean during spring and summer. Most steelhead use the ocean for two to three 1314 years before again returning to freshwater for spawning (Quinn 2005). 1315 Steelhead may be vulnerable to climate change during several portions of their life cycle 1316 that are similar in many ways to coho and Chinook salmon. Summer run adults may encounter 1317 thermally stressful temperatures during upstream migrations, which may force them to seek cold 1318 microrefugia and delay migrations (Keefer et al. 2009). Access to upstream spawning areas 1319 could be limited by ongoing declines in summer flows if passage barriers occur at road culverts 1320 or intermittency occurs in some reaches. Because summer steelhead remain for extended periods 1321 in tributaries prior to spawning, flow declines and increasing temperatures place additional 1322 stresses on these fish that may increase pre-spawn mortality rates or adversely affect their 1323 spawning ability and the viability of eggs and embryos. Juveniles of both winter and summer run 1324 fish rear for one or more years in relatively steep channels where they may be vulnerable to more 1325 frequent or larger disturbances associated with wildfires and debris flows or floods and scour 1326 (Goode et al. 2012, Sloat et al. 2017). Juveniles outmigrating through the warm downstream 1327 portions of natal rivers and the Columbia River during the spring and summer are preyed upon 1328 by the smallmouth bass (Micropterus dolomieu Lacepède, 1802) population. Like all 1329 anadromous species, interactions among climate stressors acting on multiple life stages could 1330 create negative synergies that amplify effects beyond individual life stages (Crozier et al. 2008, 1331 Honea et al. 2016). 1332 1333 1334 Bull Trout 1335 1336 Bull trout populations occur in the Lewis River and West Fork of the Klickitat River core areas 1337 (LCFRB 2010, USFWS 2015) and two natal streams with ~20 km of habitat in Pine and Rush 1338 creeks on the GPNF (table 3.7; figs. 3.10, 3.11). Bull trout may exhibit fluvial or resident life 1339 histories, but adults of both forms spawn only in the coldest headwater habitats where average 1340 summer water temperatures average less than 11 °C (Isaak et al. 2015). As a result, bull trout 1341 distributionsDRAFT are typically fragmented and confined to isolated headwater environments 32

Hudec et al. Chapter 3

1342 throughout river networks (Rieman and McIntyre 1995), although larger fish often move 1343 throughout warmer streams and rivers (Howell et al. 2010). These larger fish often predate on 1344 smaller fish or salmon eggs (Furey et al. 2017), so declines in salmon populations have probably 1345 decreased important food resources (LCFRB 2010). 1346 Bull trout spawn in the fall, and eggs incubate throughout the winter before juveniles 1347 hatch and emerge from stream substrates in late winter or early spring where rearing occurs for 1348 2-3 years (Dunham et al. 2008). Bull trout populations typically occur at low densities due to the 1349 cold and unproductive habitats that are occupied (Isaak et al. 2017b, Rieman et al. 2006). The 1350 upstream extent of bull trout habitats in the assessment area is often limited by geologic barriers 1351 that the fish are not capable of surmounting. Long-standing uncertainties about upstream fish 1352 occurrence were recently resolved by systematic surveys using environmental DNA (Carim et al. 1353 2016) that confirmed the absence of populations above barriers (Young et al. 2017). 1354 Warmer temperatures and declining summer streamflows will have negative effects on 1355 bull trout populations by reducing habitat volume and shifting suitable natal habitats further 1356 upstream if barriers are not limiting. Warmer temperatures could also facilitate the invasion of 1357 bull trout habitats by brook trout or other competitor species with warmer thermal niches that are 1358 often excluded by unsuitably cold temperatures (Isaak et al. 2015). Bull trout may be sensitive to 1359 larger or more frequent future winter high-flow events, because eggs incubate in stream 1360 substrates throughout the winter where they are susceptible to bed-scour, especially in the steeper 1361 and more confined natal habitat channels used by bull trout (Goode et al. 2013, Wenger et al. 1362 2011). If wildfires become more prevalent, they may cause more extensive geomorphic 1363 disturbances and debris flows into bull trout natal areas, given strong connectivity to adjacent 1364 hillslopes (Miller et al. 2003, Sedell et al. 2015). Negative synergies could arise by the 1365 confluence of warmer temperatures, lower flows, and increased environmental stochasticity that 1366 create larger risks than any factor individually (Jackson et al. 2009). 1367 1368 1369 Adaptation to Climate Change 1370 1371 Biological Responses 1372 1373 A changing climate and associated environmental trends will create additional stress for 1374 populations of the focal fish species in the assessment area. Habitat shifts, biotic interactions, and 1375 the potential for increased disturbances that are already depressed from more than a century of 1376 intensive land use and alterations of stream and river environments. Niche conservatism suggests 1377 there is little capacity for rapid evolutionary or physiological adaptations to warmer water 1378 temperatures (McCullough et al. 2009, Wiens et al. 2010) but trout and salmon species are 1379 noteworthy for their phenotypic plasticity, vagility, and resilience (Northcote 1992, Quinn 2005). 1380 Where barriers do not impede movements, species may adapt by shifting their 1381 distributions in space or time to track suitable habitats or to recolonize previously disturbed 1382 habitats from nearby refugia if a diversity of landscape conditions exist (Reeves et al. 1995, 1383 Sedell et al. 1990). Many of the species considered here also have diverse life histories, which 1384 may change based on how climate change affects metabolic rates, water temperature, stream 1385 productivity, and habitat connectivity. Development of adaptive responses associated with 1386 phenology may also bolster population resilience in ways that allow species to persist in dynamic 1387 environmentsDRAFT subject to long-term climate trends (Crozier et al. 2008, Kovach et al. 2012). 33

Hudec et al. Chapter 3

1388 Decreased distribution and abundance of fish populations attributable to climate change 1389 have not yet been widely documented, despite rapid climate change in the Pacific Northwest for 1390 several decades (Arismendi et al. 2013, Isaak et al. 2018, Luce and Holden 2009, Luce et al. 1391 2013). Trends toward improving freshwater habitat conditions within the assessment area 1392 streams (Lanigan et al. 2012) may also be playing a role in ameliorating potentially negative 1393 climate effects. If declines are occurring, they could be masked by a cycle of favorable ocean 1394 conditions that contributed to increased regional abundance of many anadromous species until 1395 recently (Kilduff et al. 2015), or it could be that existing monitoring programs and available 1396 datasets are inadequate for detecting the relatively subtle biological responses that are related to 1397 climate change (Crozier et al. 2011, Eby et al. 2014). Regardless, as the thermal and hydrologic 1398 changes attributable to climate warming continue to increase later this century, increasingly 1399 negative effects on cold-water fishes in Southwest Washington are expected. 1400 1401 1402 Management Responses 1403 1404 The pervasive nature of climate change will require multiple conservation strategies to bolster 1405 the resilience of individual species and populations. Conservation options will vary by location 1406 because current and future suitable habitats are expected to be more abundant and persistent in 1407 some river basins than other across the assessment area. Where habitat conditions are productive, 1408 maintaining those conditions and avoiding significant new impairments may be all that is 1409 necessary to ensure the persistence of focal species. In contrast, few habitats that function as 1410 suitable habitats may occur in other basins or where current habitats for some species are very 1411 limited. Those circumstances favor strategic, active management to promote population 1412 persistence, whether by manipulating habitat, fish populations, or both. And because many 1413 habitats are situated in landscapes that have multiple resource values and administrative 1414 agencies, balancing among competing interests will remain an underlying management theme 1415 (Rieman et al. 2010). 1416 Many things can be done to respond to climate change and improve the resilience of 1417 aquatic species as summarized in chapter 8 and in the Climate Change Adaptation Library for the 1418 Western United States (http://www.adaptationpartners.org/library.php). Adaptation options have 1419 also been the subject of several reviews, including those specific to the Pacific Northwest 1420 (Beechie et al. 2008, 2013; Halofsky and Peterson 2017; ISAB 2007; Luce et al. 2012; Rieman 1421 and Isaak 2010). Several key themes emerge from these reviews: (1) strategically prioritize and 1422 restore natural regimes of flow, sediment, wood, and temperature, (2) manage fluvial 1423 connectivity and assisted migration, (3) maintain and diversify monitoring programs, and (4) 1424 remove or suppress nonnative species. 1425 1426 Strategically prioritize restoration of natural thermal, hydrologic, and wood regimes— 1427 Future summer temperature increases and flow decreases will be the greatest stresses on cold- 1428 water fishes in the SWAP area. Improving the resilience of salmon and trout populations, 1429 therefore, may often be achieved by employing proven habitat restoration techniques that help 1430 restore natural regimes and landscape conditions. Direct solar insolation is the biggest factor 1431 contributing to stream heating (Johnson 2003, Webb and Zhang 1997), so protecting and 1432 enhancing riparian areas to maximize shade is an obvious action. In smaller streams and rivers 1433 where riparianDRAFT conditions are significantly degraded, fully functional riparian vegetation 34

Hudec et al. Chapter 3

1434 communities could offset most future stream temperature increases (Johnson and Wilby 2015, 1435 Nusslé et al. 2015), although the effectiveness of this tactic decreases in large rivers (Cristea and 1436 Burges 2010). Increased shade could be achieved by decommissioning or relocating roads away 1437 from streams, as well as grazing reductions or exclusion of livestock that may also result in 1438 stronger banks and root masses that help narrow unnaturally wide channels over time (Naiman et 1439 al. 2010). 1440 As riparian areas recover, they will provide large woody debris for channels that help 1441 diversify habitats, which would also increase channel roughness and force more instream water 1442 exchange with cooler hyporheic flows (Arrigoni et al. 2008, Luce and Caissie 2017). 1443 Enhancements of habitat and thermal diversity might also be achieved by reconnecting rivers to 1444 floodplains (Beechie et al. 2013) or restoring populations of American beaver (Castor 1445 canadensis Kuhl) to historically occupied areas (Bouwes et al. 2016, Pollock et al. 2014). 1446 Minimizing flow diversions, especially during the thermally stressful summer period, has an 1447 important cooling effect (Elmore et al. 2015) and increases habitat volume so less crowding is 1448 experienced by territorial species like salmon and trout (Null et al. 2017). In most instances, 1449 opportunities to pursue these restoration tactics will outstrip available resources, so strategic 1450 prioritization is important to ensure work is done in the most important places (Peterson et al. 1451 2013). 1452 1453 Manage fluvial connectivity and assisted migration— 1454 Obstacles to fish movements may be removed to enhance the success of migratory life history 1455 forms and species, allow fish distributions to track shifting habitats, and permit native species to 1456 reoccupy former habitat or supplement existing populations (Chelgren and Dunham 2015). 1457 Several dams currently preclude salmon access to upstream portions of river networks in the 1458 assessment area, so significant range extensions and population increases could be achieved with 1459 that tactic. Moreover, upstream habitats are typically cooler than those in downstream areas and 1460 will increasingly serve as climate refugia that ameliorate thermal stress to those migratory fish 1461 that successful return upstream. In some instances, reopening access to historical habitats may 1462 also provide access for undesirable nonnative species so context-specific assessments are needed 1463 (Fausch et al. 2009). 1464 Assisted migration—a human-assisted colonization of fish species outside their native 1465 ranges or into previously fishless waters—is a common activity in fisheries management (Rahel 1466 2016). Numerous geological barriers and waterfalls in the SWAP area naturally fragment the 1467 drainage network and historically prevented access by fish to many streams. For a non- 1468 anadromous species of concern like bull trout, these streams could serve as high-quality long- 1469 term climate refugia (Isaak et al. 2016b) and the Climate-Shield distribution model provides 1470 maps with stream-specific probabilities of habitat occupancy for detailed assessments 1471 (https://www.fs.fed.us/rm/boise/AWAE/projects/ClimateShield.html; Isaak et al. 2015). Moving 1472 native fish to such areas is feasible but sometimes controversial, because other at-risk native taxa 1473 may be vulnerable to predation or competition with native fish species (Dunham et al. 2011, 1474 2016). Nonetheless, the option is increasingly being explored for bull trout, given successful 1475 recent reintroduction efforts (Hayes and Banish 2017) and the discovery of resident populations 1476 isolated above natural barriers in other portions of the species’ range (Young et al. 2017). 1477 1478 MaintainDRAFT and diversify aquatic monitoring programs— 35

Hudec et al. Chapter 3

1479 Habitat, climate, and biological data from monitoring programs is essential for developing better 1480 information about the status and trends of aquatic habitats and populations. Maintaining and 1481 improving these monitoring programs is essential to developing improved understanding, 1482 predictive abilities, and information that will be needed by decisionmakers in future decades. 1483 The extensive temperature monitoring datasets collected by local biologists, for example, 1484 enabled the NorWeST stream temperature model and scenarios to be calibrated specifically to 1485 streams and rivers in the assessment area for maximum accuracy. Future flow scenarios might 1486 also be enhanced with additional discharge data that are often lacking in headwater catchments. 1487 Recent advances in inexpensive and reliable flow sensors (Stamp et al. 2014) make it possible to 1488 deploy sensor arrays much the way that inexpensive temperature arrays were deployed in 1489 previous decades (Dunham et al. 2005, Isaak et al. 2013). 1490 Biological data that describe species occurrence and abundance can also be obtained 1491 rapidly and inexpensively using proven eDNA sampling protocols (Carim et al. 2016, McKelvey 1492 et al. 2016). Those data, combined with existing geospatial datasets and habitat assessments, 1493 would enable refinement of information about species distributions, habitat preferences, and 1494 sensitivity to climate change or other anthropogenic stressors. In addition, eDNA samples 1495 contain the DNA of all aquatic taxa, so could provide information about many species other than 1496 cold-water fishes that often lack data but which may become species of concern in the future. 1497 Thousands of eDNA samples are now collected annually within the Pacific Northwest through 1498 dozens of agency partnerships with the U.S. Forest Service National Genomics Center for 1499 Wildlife and Fish Conservation. A database providing access to these samples is publically 1500 available at the Aquatic eDNAtlas website 1501 (https://www.fs.fed.us/rm/boise/AWAE/projects/eDNAtlas.html; Young et al. 2018) to enhance 1502 aquatic biological monitoring and modeling in future years. 1503 1504 Detection and removal of nonnative species— 1505 Removal or suppression of nonnative species that compete with focal species may also be 1506 important for maintaining or restoring the resilience of some populations (Levin et al. 2002, 1507 Sanderson et al. 2009). Early detection of invasions is key for successful control efforts, so 1508 routine eDNA monitoring of priority habitats might be done to provide an early warning system. 1509 Once nonnative species are established, control efforts involving electrofishing removals or 1510 chemical treatments are often costly, in part because they need to be conducted on multiple 1511 occasions to be effective (Buktenica et al. 2013), and may be successful only in smaller 1512 headwater streams (Shepard et al. 2002). Furthermore, success with either method is obtained 1513 only if the source of nonnative species is removed. 1514 In headwater streams, artificial barriers are sometimes created to block upstream 1515 dispersal by invasives (Rahel 2013), a service that natural geological barriers might perform in 1516 some SWAP area streams if assisted migration options were pursued for bull trout populations. 1517 Finally, using control measures to reduce the abundance of nonnative species rather than remove 1518 them has been applied in some streams (e.g., suppression of brook trout populations by regular 1519 electrofishing to favor bull trout; Peterson et al. 2008). Such activities are likely to be successful 1520 only if conducted at regular intervals on a continual basis, which assumes funding and 1521 enthusiasm for such ventures will be available indefinitely. 1522 1523 SummaryDRAFT— 36

Hudec et al. Chapter 3

1524 Responding to the environmental trends associated with climate change will require a diverse 1525 portfolio comprised of many of the actions described above. Equally important will be adapting 1526 our mindsets and administrative processes to a new paradigm of dynamic disequilibrium for the 1527 21st century. Stream habitats will become more variable, undergo gradual shifts through time, 1528 and sometimes decline. Many populations are resilient enough to persist in, or track, suitable 1529 habitats, but others could be overwhelmed by future changes. It is unlikely that we will be able to 1530 preserve all populations of aquatic species as they currently exist this century. However, as better 1531 monitoring programs and information continue to be developed, managers will have ever more 1532 precise tools at their disposal to know when and where resource commitments are best made to 1533 enhance the resilience of existing populations or to benefit other species for which management 1534 was previously not a priority. 1535 1536 1537 Acknowledgments 1538 1539 We thank the many biologists who contributed ideas, refinements, and data to the assessment 1540 described in this chapter and who participated in the workshop Vancouver, WA. Comments 1541 provided by Sam Brenkman, Correigh Greene, and Karl Polivka improved the quality of the final 1542 assessment. 1543 1544 1545 Literature Cited 1546 1547 Al-Chokhachy, R.; Schmetterling, D.; Clancy, C. [et al.]. 2016. Are brown trout replacing or 1548 displacing bull trout populations in a changing climate? Canadian Journal of Fisheries and 1549 Aquatic Sciences. 73: 1395–1404. 1550 Arismendi, I.; Safeeq, M.; Johnson, S.L. [et al.]. 2013. Increasing synchrony of high 1551 temperature and low flow in western North American streams: double trouble for coldwater 1552 biota? Hydrobiologia. 712: 61–70. 1553 Arrigoni, A.S.; Poole, G.C.; Mertes, L.A. [et al.]. 2008. Buffered, lagged, or cooled? 1554 Disentangling hyporheic influences on temperature cycles in stream channels. Water 1555 Resources Research. 44: W09418. 1556 Baigun, C.R. 2003. Characteristics of deep pools used by adult summer steelhead in Steamboat 1557 Creek, Oregon. North American Journal of Fisheries Management. 23: 1167–1174. 1558 Bartholow, J.M. 2005. Recent water temperature trends in the lower Klamath River, California. 1559 North American Journal of Fish Management. 25: 152–162. 1560 Beamish, R.J.; Mahnken, C. 2001. A critical size and period hypothesis to explain natural 1561 regulation of salmon abundance and the linkage to climate and climate change. Progress in 1562 Oceanography. 49: 423–437. 1563 Beechie, T.; Imaki, H.; Greene, J. [et al.]. 2013. Restoring salmon habitat for a changing 1564 climate. River Research and Applications. 29: 939–960. 1565 Beechie, T.; Pess, G.; Roni, P.; Giannico, G. 2008. Setting river restoration priorities: a review 1566 of approaches and a general protocol for identifying and prioritizing actions. North American 1567 Journal of Fisheries Management. 28: 891–905. 1568 Beschta, R.L. 1978. Long-term patterns of sediment production following road construction and 1569 loggingDRAFT in the Oregon Coast Range. Water Resources Research. 14: 1011–1016. 37

Hudec et al. Chapter 3

1570 Bisson, P.A.; Rieman, B.E.; Luce, C. [et al.]. 2003. Fire and aquatic ecosystems of the western 1571 USA: current knowledge and key questions. Forest Ecology and Management. 178: 213–229. 1572 Bond, N.A.; Cronin, M.F.; Freeland, H.; Mantua, N. 2015. Causes and impacts of the 2014 1573 warm anomaly in the NE Pacific. Geophysical Research Letters. 42: 3414–3420. 1574 Bouwes, N.; Weber, N.; Jordan, C.E. [et al.]. 2016. Ecosystem experiment reveals benefits of 1575 natural and simulated beaver dams to a threatened population of steelhead (Oncorhynchus 1576 mykiss). Scientific Reports. 6: 28581. 1577 Bowerman, T.; Keefer, M.L.; Caudill, C.C. 2016. Pacific salmon prespawn mortality: patterns, 1578 methods, and study design considerations. Fisheries. 41: 738–749. 1579 Buktenica, M.W.; Hering, D.K.; Girdner, S.F. [et al.]. 2013. Eradication of nonnative brook 1580 trout with electrofishing and Antimycin-A and the response of a remnant bull trout 1581 population. North American Journal of Fisheries Management. 33: 117–129. 1582 Burnett, K.M.; Reeves, G.H.; Miller, D.J. [et al.]. 2007. Distribution of salmon-habitat 1583 potential relative to landscape characteristics and implications for conservation. Ecological 1584 Applications. 17: 66–80. 1585 Carim, K.J.; McKelvey, K.S.; Young, M.K. [et al.]. 2016. A protocol for collecting 1586 environmental DNA samples from streams. Gen. Tech. Rep. RMRS-GTR-355. Fort Collins, 1587 CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station. 1588 Chelgren, N.D.; Dunham, J.B. 2015. Connectivity and conditional models of access and 1589 abundance of species in stream networks. Ecological Applications. 25: 1357–1372. 1590 Comte, L.; Buisson, L.; Daufresne, M.; Grenouillet, G. 2013. Climate-induced changes in the 1591 distribution of freshwater fish: observed and predicted trends. Freshwater Biology. 58: 625– 1592 639. 1593 Cooke, S.J.; Hinch, S.G.; Farrell, A.P. [et al.]. 2004. Abnormal migration timing and high en 1594 route mortality of sockeye salmon in the Fraser River, British Columbia. Fisheries. 29: 22– 1595 33. 1596 Cristea, N.C.; Burges, S.J. 2010. An assessment of the current and future thermal regimes of 1597 three streams located in the Wenatchee River basin, Washington State: Some implications for 1598 regional river basin systems. Climatic Change. 102: 493–520. 1599 Crozier, L.G.; Hendry, A.P.; Lawson, P.W. [et al.]. 2008. Potential responses to climate 1600 change in organisms with complex life histories: evolution and plasticity in Pacific salmon. 1601 Evolutionary Applications. 1: 252–270. 1602 Crozier L.G.; Scheuerell, M.D.; Zabel, R.W. 2011. Using time series analysis to characterize 1603 evolutionary and plastic responses to environmental change: a case study of a shift toward 1604 earlier migration date in sockeye salmon. The American Naturalist. 178: 755–773. 1605 Di Lorenzo, E.; Mantua, N. 2016. Multi-year persistence of the 2014/15 North Pacific marine 1606 heatwave. Nature Climate Change. 6: 1042–1047. 1607 Dose, J.J.; Roper, B.B. 1994. Long-term changes in low-flow channel widths within the South 1608 Umpqua watershed, Oregon. Journal of the American Water Resources Association. 30: 993– 1609 1000. 1610 Dunham, J.; Baxter, C.; Fausch, K. [et al.]. 2008. Evolution, ecology, and conservation of 1611 Dolly Varden, white spotted char, and bull trout. Fisheries. 33: 537–550. 1612 Dunham, J.B.; Chandler G.; Rieman B.E.; Martin, D. 2005. Measuring stream temperature 1613 with digital dataloggers: a user’s guide. Gen. Tech. Rep. RMRS-GTR-150WWW. Fort 1614 Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 1615 Station.DRAFT 38

Hudec et al. Chapter 3

1616 Dunham, J.B.; Gallo, K.; Shively, D. [et al.]. 2011. Assessing the feasibility of native fish 1617 reintroductions: a framework applied to threatened bull trout. North American Journal of 1618 Fisheries Management. 31: 106–115. 1619 Dunham, J.B.; White, R.; Allen, C.S. [et al.]. 2016. The reintroduction landscape. In: 1620 Jachowski, D.S.; Millspaugh, J.J.; Angermeier, P.L.; Slotow, R., eds. Reintroduction of fish 1621 and wildlife populations. Berkeley, CA: University of California Press: 79–103. 1622 Dunham, J.B.; Young, M.K.; Gresswell, R.E.; Rieman, B.E. 2003. Effects of fire on fish 1623 populations: landscape perspectives on persistence of native fishes and non-native fish 1624 invasions. Forest Ecology and Management. 178: 183–196. 1625 Ebersole, J.L.; Liss, W.J.; Frissell, C.A. 2001. Relationship between stream temperature, 1626 thermal refugia and rainbow trout Oncorhynchus mykiss abundance in arid-land streams in 1627 the northwestern United States. Ecology of Freshwater Fish. 10: 1–10. 1628 Eby, L.A.; Helmy, O.; Holsinger, L.M.; Young, M.K. 2014. Evidence of climate-induced 1629 range contractions for bull trout in a Rocky Mountain watershed, U.S.A. PLoS ONE 9: 1630 e98812. 1631 Elmore, L.R.; Null, S.E.; Mouzon, N.R. 2015. Effects of environmental water transfers on 1632 stream temperatures. River Research and Applications. 32: 1415–1427. 1633 Fausch, K.D.; Rieman, B.E.; Dunham, J.B.; Young, M.K.; Peterson, D.P. 2009. Invasion 1634 versus isolation: Trade-offs in managing native salmonids with barriers to upstream 1635 movement. Conservation Biology. 23: 859–870. 1636 Ford, M.J.; Cooney, T.; McElhany, P. [et al.]. 2011. Status review update for Pacific salmon 1637 and steelhead listed under the Endangered Species Act: Pacific Northwest. NOAA Tech. 1638 Memorand. NMFS-NWFSC-113. Seattle, WA: U.S. Department of Commerce, National 1639 Oceanic and Atmospheric Administration, Northwest Fisheries Science Center. 1640 Furey, N.; Hinch, S.G. 2017. Bull trout movements match the life history of sockeye salmon: 1641 consumers can exploit seasonally distinct resource pulses. Transactions of the American 1642 Fisheries Society. 146: 450–461. 1643 Furniss, M.J.; Roby, K.B.; Cenderelli, D. [et al.]. 2013. Assessing the vulnerability of 1644 watersheds to climate change: results of national forest watershed vulnerability pilot 1645 assessments. Gen. Tech. Rep. PNW-GTR-884. Portland, OR: U.S. Department of 1646 Agriculture, Forest Service, Pacific Northwest Research Station. 1647 Furniss, M.J.; Staab, B.P.; Hazelhurst, S. [et al.]. 2010. Water, climate change, and forests: 1648 watershed stewardship for a changing climate. Gen. Tech. Rep. PNW-GTR-812. Portland, 1649 OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 1650 Goode, J.R.; Buffington, J.M.; Tonina, D. [et al.]. 2013. Potential effects of climate change on 1651 streambed scour and risks to salmonid survival in snow-dominated mountain basins. 1652 Hydrological Processes. 27: 750–765. 1653 Goode, J.R.; Luce, C.H.; Buffington, J.M. 2012. Enhanced sediment delivery in a changing 1654 climate in semi-arid mountain basins: Implications for water resource management and 1655 aquatic habitat in the northern Rocky Mountains. Geomorphology. 139–140: 1–15. 1656 Halofsky, J.E.; Peterson, D.L., eds. 2017. Climate change vulnerability and adaptation in the 1657 Blue Mountains. Gen. Tech. Rep. PNW-GTR-939. Portland, OR: U.S. Department of 1658 Agriculture, Forest Service, Pacific Northwest Research Station. 1659 Hamlet, A.F.; Lettenmaier, D.P. 2007. Effects of 20th century warming and climate variability 1660 on floodDRAFT risk in the western US. Water Resources Research. 43: W06427. 39

Hudec et al. Chapter 3

1661 Hamlet, A.F.; Mote, P.W.; Clark, M.P.; Lettenmaier, D.P. 2007. 20th century trends in runoff, 1662 evapotranspiration, and soil moisture in the Western U.S. Journal of Climate. 20: 1468–1486. 1663 Hare, S.R.; Mantua N.J.; Francis R.C. 1999. Inverse production regimes: Alaska and West 1664 Coast Pacific salmon. Fisheries. 24: 6–14. 1665 Hauer, F.R.; Baron, J.S.; Campbell, D.H. [et al.]. 1997. Assessment of climate change and 1666 freshwater ecosystems of the Rocky Mountains, USA and Canada. Hydrological Processes. 1667 11: 903–924. 1668 Hayes, M.F.; Banish, N.P. 2017. Translocation and reintroduction of native fishes: a review of 1669 bull trout Salvelinus confluentus with applications for future reintroductions. Endangered 1670 Species Research. 34: 191–209. 1671 Healey, M.C. 1991. Life history of chinook salmon (Oncorhynchus tshawytscha). In: Groot, C; 1672 Margolis, L., eds. Pacific salmon life histories. Vancouver, British Columbia, Canada: 1673 University of British Columbia Press: 313–393. 1674 Hessburg, P.F.; Agee, J.K. 2003. An environmental narrative on inland northwest United States 1675 forests, 1800-2000. Forest Ecology and Management. 178: 23–59. 1676 Holtby, L.B. 1988. Effects of logging on stream temperatures in Carnation Creek British 1677 Columbia, and associated impacts on the coho salmon (Oncorhynchus kisutch). Canadian 1678 Journal of Fisheries and Aquatic Sciences. 45: 502–515. 1679 Honea, J.M.; McClure, M.M.; Jorgensen, J.C.; Scheuerell, M.D. 2016. Assessing freshwater 1680 life-stage vulnerability of an endangered Chinook salmon population to climate change 1681 influences on stream habitat. Climate Research. 71: 127–137. 1682 Hotaling, S.; Finn, D.S.; Giersch, J.J. [et al.]. 2017. Climate change and alpine stream biology: 1683 progress, challenges, and opportunities for the future. Biological Reviews. 92: 2024–2045. 1684 Howell, P.J.; Dunham, J.B.; Sankovich, P.M. 2010. Relationships between water temperatures 1685 and upstream migration, cold water refuge use, and spawning of adult bull trout from the 1686 Lostine River, Oregon, USA. Ecology of Freshwater Fish. 19: 96–106. 1687 Independent Science Advisory Board [ISAB]. 2007. Climate change impacts on Columbia 1688 River basin fish and wildlife. ISAB Clim. Change Rep. ISAB 2007-2. Portland, OR: 1689 Northwest Power and Conservation Council. 1690 Isaak, D.J.; Horan, D.L.; Wollrab, S.P. 2013. A simple protocol using underwater epoxy to 1691 install annual temperature monitoring sites in rivers and streams. Gen. Tech. Rep. RMRS- 1692 GTR-314. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 1693 Mountain Research Station. 1694 Isaak, D.J.; Luce C.H.; Horan, D.L. [et al.]. 2018. Global warming of salmon and trout rivers 1695 in the Northwestern U.S.: road to ruin or path through purgatory? Transactions of the 1696 American Fisheries Society. In press. 1697 Isaak, D.J.; Wenger, S.J.; Peterson, E.E. [et al.]. 2016a. NorWeST modeled summer stream 1698 temperature scenarios for the western U.S. Research data archive. Fort Collins, CO: U.S. 1699 Department of Agriculture, Forest Service, Rocky Mountain Research Station, 1700 https://doi.org/10.2737/RDS-2016-0033. (21 December 2017). 1701 Isaak, D.J.; Wenger S.J.; Peterson E.E. [et al.]. 2017a. The NorWeST summer temperature 1702 model and scenarios for the western U.S.: a crowd-sourced database and new geospatial tools 1703 foster a user-community and predict broad climate warming of rivers and streams. Water 1704 ResourcesDRAFT Research. 53: 9181–9205. 40

Hudec et al. Chapter 3

1705 Isaak, D.J.; Wenger, S.J.; Young, M.K. 2017b. Big biology meets microclimatology: defining 1706 thermal niches of ectotherms at landscape scales for conservation planning. Ecological 1707 Applications. 27: 977–990. 1708 Isaak, D.J.; Wollrab, S.; Horan, D.; Chandler, G. 2012. Climate change effects on stream and 1709 river temperatures across the northwest U.S. from 1980–2009 and implications for salmonid 1710 fishes. Climatic Change. 113: 499–524. 1711 Isaak, D.J.; Young, M.K.; Luce, C. [et al.]. 2016b. Slow climate velocities of mountain 1712 streams portend their role as refugia for cold-water biodiversity. Proceedings of the National 1713 Academy of Sciences, USA. 113: 4374–4379. 1714 Isaak, D.J.; Young, M.K.; Nagel, D. [et al.]. 2015. The cold-water climate shield: delineating 1715 refugia to preserve salmonid fishes through the 21st century. Global Change Biology. 21: 1716 2540–2553. 1717 Jackson, S.T.; Betancourt, J.L.; Booth, R.K., Gray, S.T. 2009. Ecology and the ratchet of 1718 events: climate variability, niche dimensions, and species distributions. Proceedings of the 1719 National Academy of Sciences, USA. 106: 19685–19692. 1720 Johnson, M.F.; Wilby, R.L. 2015. Seeing the landscape for the trees: metrics to guide riparian 1721 shade management in river catchments. Water Resources Research. 51: 3754–3769. 1722 Johnson, S.L. 2003. Stream temperature: scaling of observations and issues for modeling. 1723 Hydrological Processes. 17: 497–499. 1724 Johnson, S.L.; Jones, J.A. 2000. Stream temperature responses to forest harvest and debris 1725 flows in western Cascades, Oregon. Canadian Journal of Fisheries and Aquatic Sciences. 1726 57(S2): 30–39. 1727 Jones, J.A.; Grant, G.E. 1996. Peak flow responses to clear-cutting and roads in small and large 1728 basins, western Cascades, Oregon. Water Resources Research. 32: 959–974. 1729 Keefer, M.L.; Peery, C.A.; High, B. 2009. Behavioral thermoregulation and associated 1730 mortality trade-offs in migrating adult steelhead (Oncorhynchus mykiss): variability among 1731 sympatric populations. Canadian Journal of Fisheries and Aquatic Science. 66: 1734–1747. 1732 Kiffney, P.M.; Bull, J.P.; Feller, M.C. 2002. Climatic and hydrologic variability in a coastal 1733 watershed of southwestern British Columbia. Journal of the American Water Resources 1734 Association. 38: 1437–1451. 1735 Kilduff, D.P.; Di Lorenzo, E.; Botsford, L.W.; Teo, S.L.H. 2015. Changing central Pacific El 1736 Niños reduce stability of North American salmon survival rates. Proceedings of the National 1737 Academy of Sciences, USA. 112: 10962–10966. 1738 Kovach, R.P.; Gharrett, A.J.; Tallmon, D.A. 2012. Genetic change for earlier migration 1739 timing in a pink salmon population. Proceedings of the Royal Society of London B: 1740 Biological Sciences. 279: 3870–3878. 1741 Lanigan, S.H.; Gordon, S.N.; Eldred, P. [et al.]. 2012. Northwest Forest Plan—the first 15 1742 years (1994-2008): watershed condition status and trend. Gen. Tech. Rep. PNW-GTR-856. 1743 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 1744 Station. 1745 Levin, P.S.; Achord, S.; Feist, B.E.; Zabel, R.W. 2002. Non-indigenous brook trout and the 1746 demise of Pacific salmon: a forgotten threat? Proceedings of the Royal Society of London B. 1747 269: 1663–1670. 1748 Lower Columbia Fish Recovery Board [LCFRB]. 2010. Washington Lower Columbia salmon 1749 recovery and fish and wildlife subbasin plan. Olympia, WA: Lower Columbia Fish Recovery 1750 Board.DRAFT 41

Hudec et al. Chapter 3

1751 Luce, C.H.; Abatzoglou, J.T.; Holden, Z.A. 2013. The missing mountain water: slower 1752 westerlies decrease orographic precipitation. Science. 266: 776–779. 1753 Luce, C.H.; Holden, Z.A. 2009. Declining annual streamflow distributions in the Pacific 1754 Northwest United States, 1948-2006. Geophysical Research Letters. 36: L16401. 1755 Luce, C.; Morgan, P.; Dwire, K. [et al.]. 2012. Climate change, forests, fire, water, and fish: 1756 Building resilient landscapes, streams, and managers. Gen. Tech. Rep. RMRS-GTR-290. 1757 Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 1758 Station. 1759 Luce, C.H.; Staab, B.P.; Kramer, M.G. 2014. Sensitivity of summer stream temperatures to 1760 climate variability in the Pacific Northwest. Water Resources Research. 50: 3428–3443. 1761 Lynch, A.J.; Myers, B.J.; Chu, C., [et al.]. 2016. Climate change effects on North American 1762 inland fish populations and assemblages. Fisheries. 41: 346–361. 1763 Lynch, D.D.; Risley, J.C. 2003, Klamath River Basin hydrologic conditions prior to the 1764 September 2002 die-off of salmon and steelhead: USGS Water Resour. Invest. Rep. 03– 1765 4099. Reston, VA: U.S. Geological Survey. 1766 Manhard, C.V.; Joyce, J.E.; Gharrett. A.J. 2017. Evolution of phenology in a salmonid 1767 population: a potential adaptive response to climate change. Canadian Journal of Fisheries 1768 and Aquatic Sciences. 74: 1519–1527. 1769 Mantua, N.J.; Hare, S.R.; Zhang, Y. [et al.]. 1997. A Pacific interdecadal climate oscillation 1770 with impacts on salmon production. Bulletin of the American Meteorological Society. 78: 1771 1069–1079. 1772 Mantua, N.J.; Tohver, I.; Hamlet, A. 2010. Climate change impacts on streamflow extremes 1773 and summertime stream temperature and their possible consequences for freshwater salmon 1774 habitat in Washington state. Climatic Change. 102: 187–223. 1775 Marlier, M.E.; Xiao, M.; Engel, R. [et al.]. 2017. The 2015 drought in Washington State: a 1776 harbinger of things to come? Environmental Research Letters. 12: 114008. 1777 Martins, E.G.; Hinch, S.G.; Cooke, S.J.; Patterson, D.A. 2012. Climate effects on growth, 1778 phenology, and survival of sockeye salmon (Oncorhynchus nerka): a synthesis of the current 1779 state of knowledge and future research directions. Reviews in Fish Biology and Fisheries. 22: 1780 887–914. 1781 May, C.L.; Gresswell, R.E. 2004. Spatial and temporal patterns of debris-flow deposition in the 1782 Oregon Coast Range, USA. Geomorphology. 57: 135–149. 1783 McCullough, D.A.; Bartholow, J.M.; Jager, H.I. [et al.]. 2009. Research in thermal biology: 1784 burning questions for coldwater stream fishes. Reviews in Fisheries Science. 17: 90–115. 1785 McKay, L.; Bondelid, T.; Dewald, T. [et al.]. 2012. NHDPlus version 2: user guide. 1786 ftp://ftp.horizon- 1787 systems.com/NHDPlus/NHDPlusV21/Documentation/NHDPlusV2_User_Guide.pdf. (21 1788 December 2017). 1789 McKean, J.; Tonina, D. 2013. Bed stability in unconfined gravel bed mountain streams: With 1790 implications for salmon spawning viability in future climates. Journal of Geophysical 1791 Research: Earth Surface. 118: 1227–1240. 1792 McKelvey, K.S.; Young, M.K.; Knotek, W.L. [et al.]. 2016. Sampling large geographic areas 1793 for rare species using environmental DNA: a study of bull trout occupancy in western 1794 Montana. Journal of Fish Biology. 88: 1215–1222. 1795 Miller, D.; Luce, C.; Benda, L. 2003. Time, space, and episodicity of physical disturbance in 1796 streams.DRAFT Forest Ecology and Management. 178: 121–140. 42

Hudec et al. Chapter 3

1797 Miller, R.R. 2010. Is the past present? Historical splash dam mapping and stream disturbance 1798 detection in the Oregon coast range, Corvallis. Master’s thesis. Corvallis OR: Oregon State 1799 University. 1800 Moore, R.; Spittlehouse, D.L.; Story, A. 2005. Riparian microclimate and stream temperature 1801 response to forest harvesting: a review. Journal of the American Water Resources 1802 Association. 41: 813–834. 1803 Moore, R.; Wondzell, S.M. 2005. Physical hydrology and the effects of forest harvesting in the 1804 Pacific Northwest: a review. Journal of the American Water Resources Association. 41: 763– 1805 784. 1806 Mote, P.W.; Parson, E.A.; Hamlet, A.F. [et al.] 2003. Preparing for climatic change: the water, 1807 salmon, and forests of the Pacific Northwest. Climatic Change. 61: 45–88. 1808 Naiman, R.J.; Décamps, H.; McClain, M.E. 2010. Riparia: ecology, conservation, and 1809 management of streamside communities. Cambridge, MA: Academic Press. 1810 National Oceanic and Atmospheric Administration [NOAA]. 2016. 2016 5-year review: 1811 summary and evaluation of Lower Columbia River Chinook salmon, Columbia River chum 1812 salmon, Lower Columbia River coho salmon, Lower Columbia River steelhead. Portland, 1813 OR: National Oceanic and Atmospheric Administration, National Marine Fisheries Service, 1814 West Coast Region. 1815 Nehlsen, W.; Williams, J.E.; Lichatowich, J.A. 1991. Pacific salmon at the crossroads: stocks 1816 at risk from California, Oregon, Idaho, and Washington. Fisheries. 16: 4–21. 1817 Nickelson, T.E.; Rodgers, J.D.; Johnson, S.L.; Solazzi, M.F. 1992. Seasonal changes in 1818 habitat use by juvenile coho salmon (Oncorhynchus kisutch) in Oregon coastal streams. 1819 Canadian Journal of Fisheries and Aquatic Sciences. 49: 783–789. 1820 Northcote, T.G. 1992. Migration and residency in stream salmonids: some ecological 1821 considerations and evolutionary consequences. Nordic Journal of Freshwater Research. 67: 1822 5–17. 1823 Northwest Fisheries Science Center [NWFSC]. 2015. Status review update for Pacific salmon 1824 and steelhead listed under the Endangered Species Act: Pacific Northwest. Seattle, WA: 1825 National Oceanic and Atmospheric Administration, Northwest Fisheries Science Center. 1826 Null, S.E.; Mouzon, N.R.; Elmore, L.R. 2017. Dissolved oxygen, stream temperature, and fish 1827 habitat response to environmental water purchases. Journal of Environmental Management. 1828 197: 559–570. 1829 Nusslé, S.; Matthews, K.R.; Carlson S.M. 2015. Mediating water temperature increases due to 1830 livestock and global change in high elevation meadow streams of the Golden Trout 1831 Wilderness. PLoS One. 10: e0142426. 1832 Oregon Department of Fish and Wildlife [ODFW]. 2015. News release: ODFW takes action 1833 to help native fish. http://dfw.state.or.us/news/2015/july/071615.asp. (21 December 2017). 1834 Petersen, J.H.; Kitchell, J.F. 2001. Climate regimes and water temperature changes in the 1835 Columbia River: bioenergetic implications for predators of juvenile salmon. Canadian 1836 Journal of Fisheries and Aquatic Sciences. 58: 1831–1841. 1837 Peterson, D.P.; Fausch, K.D.; Watmough, J.; Cunjak, R.A. 2008. When eradication is not an 1838 option: modeling strategies for electrofishing suppression of nonnative brook trout to foster 1839 persistence of sympatric native cutthroat trout in small streams. North American Journal of 1840 Fisheries Management. 28: 1847–1867. 1841 Pollock, M.M.; Beechie, T.J.; Wheaton, J.M. [et al.]. 2014. Using beaver dams to restore 1842 incisedDRAFT stream ecosystems. BioScience. 64: 279–290. 43

Hudec et al. Chapter 3

1843 Quinn, T.P. 2005. The behavior and ecology of Pacific salmon and trout. Seattle, WA: 1844 University of Washington Press. 1845 Rahel, F.J., 2013. Intentional fragmentation as a management strategy in aquatic 1846 systems. BioScience. 63: 362–372. 1847 Rahel, F.J. 2016. Changing philosophies of fisheries management as illustrated by the history of 1848 fishing regulations in Wyoming. Fisheries. 41: 38–48. 1849 Ratner, S.; Lande, R.; Roper B.B. 1997. Population viability analysis of spring Chinook 1850 salmon in the South Umpqua River, Oregon. Conservation Biology. 11: 879–889. 1851 Reeves, G.H.; Benda, L.E.; Burnett, K.M. [et al.]. 1995. A disturbance-based ecosystem 1852 approach to maintaining and restoring freshwater habitats of evolutionarily significant units 1853 of anadromous salmonids in the Pacific Northwest. In: American Fisheries Society 1854 Symposium 17. Bethesda, MD: American Fisheries Society: 334–349. 1855 Reeves, G.H.; Bisson P. A.; Dambacher J.M. 1998. Fish communities. In: Naiman, R.J.; Bilby, 1856 R.E., eds. River ecology and management: lessons from the Pacific coastal ecoregion. New 1857 York, NY: Springer-Verlag: 200–234. 1858 Reeves, G.H.; Hohler, D.B.; Larsen, D.P. [et al.]. 2003. Effectiveness monitoring for the 1859 aquatic and riparian component of the Northwest Forest Plan: conceptual framework and 1860 options. Gen. Tech. Rep. PNW-GTR-577. Portland, OR: U.S. Department of Agriculture, 1861 Forest Service, Pacific Northwest Research Station. 1862 Reeves, G.H.; Olson, D.H.; Wondzell, S.M. [et al]. 2018. The aquatic conservation strategy of 1863 the Northwest Forest Plan—a review of the relevant science after 22 years. In: Spies, T.; 1864 Stine, P.; Gravenmier, R. [et al.], eds. Synthesis of science to inform land management 1865 within the Northwest Forest Plan area. Gen. Tech. Rep. PNW-GTR-xxx. Portland, OR: U.S. 1866 Department of Agriculture, Forest Service, Pacific Northwest Research Station. In press. 1867 Rieman, B.E.; Beamesderfer, R.C.; Vigg, S.; Poe, T.P. 1991. Estimated loss of juvenile 1868 salmonids to predation by northern squawfish, walleyes, and smallmouth bass in John Day 1869 Reservoir, Columbia River. Transactions of the American Fisheries Society. 120: 448–458. 1870 Rieman, B.E.; Hessburg, P.F.; Luce, C.; Dare, M.R. 2010. Wildfire and management of 1871 forests and native fishes: conflict or opportunity for convergent solutions? BioScience. 60: 1872 460–468. 1873 Rieman, B.E.; Isaak, D.J. 2010. Climate change, aquatic ecosystems and fishes in the Rocky 1874 Mountain West: implications and alternatives for management. Gen. Tech. Rep. GTR- 1875 RMRS-250. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 1876 Mountain Research Station. 1877 Rieman, B.E.; McIntyre, J.D. 1995. Occurrence of bull trout in naturally fragmented habitat 1878 patches of varied size. Transactions of the American Fisheries Society. 124: 285–296. 1879 Rieman, B.E.; Peterson, J.T.; Myers, D.L. 2006. Have brook trout (Salvelinus fontinalis) 1880 displaced bull trout (Salvelinus confluentus) along longitudinal gradients in central Idaho 1881 streams? Canadian Journal of Fisheries and Aquatic Sciences. 63: 63–78. 1882 Sanderson, B.L.; Barnas, K.A.; Wargo-Rub A.M. 2009. Nonindigenous species of the Pacific 1883 Northwest: an overlooked risk to endangered salmon? BioScience. 59: 245–256. 1884 Schindler, D.E.; Hilborn, R.; Chasco, B. [et al.]. 2010. Population diversity and the portfolio 1885 effect in an exploited species. Nature. 465: 609–612. 1886 Sedell, J.R.; Froggatt J.L. 1984. Importance of streamside forests to large rivers: the isolation 1887 of the Willamette River, Oregon, USA, from its floodplain by snagging and streamside forest 1888 removal.DRAFT Verhandlungen des Internationalen Verein Limnologie. 22: 1828–1834. 44

Hudec et al. Chapter 3

1889 Sedell, E.R.; Gresswell, R.E.; McMahon, T.E. 2015. Predicting spatial distribution of postfire 1890 debris flows and potential consequences for native trout in headwater streams. Freshwater 1891 Science. 34: 1558–1570. 1892 Sedell, J.R.; Reeves, G.H.; Hauer, F.R. [et al.]. 1990. Role of refugia in recovery from 1893 disturbances: modern fragmented and disconnected river systems. Environmental 1894 Management. 14: 711–724. 1895 Shellberg, J.G.; Bolton, S.B.; Montgomery, D.R. 2010. Hydrogeomorphic effects on bedload 1896 scour in bull char (Salvelinus confluentus) spawning habitat, western Washington, USA. 1897 Canadian Journal of Fisheries and Aquatic Sciences. 67: 626–640. 1898 Shepard, B.B.; Spoon, R.; Nelson, L. 2002. A native westslope cutthroat trout population 1899 responds positively after brook trout removal and habitat restoration. Intermountain Journal 1900 of Sciences. 8: 193–214. 1901 Sloat, M.R.; Reeves, G.H.; Christiansen, K.R. 2017. Stream network geomorphology mediates 1902 predicted vulnerability of anadromous fish habitat to hydrologic change in southeast Alaska. 1903 Global Change Biology. 23: 604–620. 1904 Stamp, J.; Hamilton A.; Craddock M. [et al.]. 2014. Best practices for continuous monitoring 1905 of temperature and flow in wadeable streams. EPA/600/R-13/170F. Washington, DC: U.S. 1906 Global Change Research Program, National Center for Environmental Assessment. 1907 Steel, E.A.; Muldoon, A.; Flitcroft, R.L. [et al.]. 2016. Current landscapes and legacies of 1908 land-use past: understanding the distribution of juvenile coho salmon (Oncorhynchus kisutch) 1909 and their habitats along the Oregon Coast, USA. Canadian Journal of Fisheries and Aquatic 1910 Sciences. 74: 546–561. 1911 Steel, E.A.; Feist, B.E.; Jensen, D.W. [et al.]. 2004. Landscape models to understand steelhead 1912 (Oncorhynchus mykiss) distribution and help prioritize barrier removals in the Willamette 1913 basin, Oregon, USA. Canadian Journal of Fisheries and Aquatic Sciences. 61: 999–1011. 1914 Torgersen, C.E.; Price D.M.; Li, H.W.; McIntosh B.A. 1999. Multiscale thermal refugia and 1915 stream habitat associations of Chinook salmon in northeastern Oregon. Ecological 1916 Applications. 9: 301–319. 1917 Trombulak, S.C.; Frissell, C.A. 2000. Review of ecological effects of roads on terrestrial and 1918 aquatic communities. Conservation Biology. 14: 18–30. 1919 U.S. Department of Agriculture, Forest Service [USDA FS]. 2005. Pacific Northwest Region 1920 aquatic conservation strategy. Portland, OR: U.S. Department of Agriculture, Forest Service, 1921 Pacific Northwest Region. 1922 U.S. Department of Agriculture, Forest Service [USDA FS]. 2011. Watershed condition 1923 framework. FS-977. Washington, DC: U.S. Department of Agriculture, Forest Service. 1924 U.S. Fish and Wildlife Service [USFWS]. 2015. Recovery plan for the coterminous United 1925 States population of bull trout (Salvelinus confluentus). Portland, OR. 1926 Webb, B.W.; Zhang Y. 1997. Spatial and seasonal variability in the components of the river 1927 heat budget. Hydrological Processes. 11: 79–101 1928 Wenger, S.J.; Isaak, D.J.; Dunham, J.B. [et al.]. 2011. Role of climate and invasive species in 1929 structuring trout distributions in the interior Columbia River Basin, USA. Canadian Journal 1930 of Fisheries and Aquatic Sciences. 68: 988–1008. 1931 Wenger, S.J.; Luce, C.H.; Hamlet, A.F. [et al.]. 2010. Macroscale hydrologic modeling of 1932 ecologically relevant flow metrics. Water Resources Research. 46: W09513. 1933 Whitney, J.E.; Al-Chokhachy, R.K.; Bunnell, D.B. [et al.]. 2016. Physiological basis of 1934 climateDRAFT change impacts on North American inland fishes. Fisheries. 41: 332–345. 45

Hudec et al. Chapter 3

1935 Wiens, J.J.; Ackerly, D.D.; Allen, A.P. [et al.]. 2010. Niche conservatism as an emerging 1936 principle in ecology and conservation biology. Ecology Letters. 13: 1310–1324. 1937 Willis, R.A. 1954. The length of time that silver salmon spent before death on spawning grounds 1938 at Spring Creek, Wilson River in 1951-1952. Fish Commission of Oregon Research Briefs. 5: 1939 27–31. 1940 Wing, M.C.; Skaugset, A. 2002. Relationships of channel characteristics, land ownership, and 1941 land use patterns to large woody debris in western Oregon streams. Canadian Journal of 1942 Fisheries and Aquatic Sciences. 59: 196–807. 1943 Young, M.K.; Isaak, D.J.; McKelvey, K.S. [et al.]. 2017. Species occurrence data from the 1944 range-wide bull trout eDNA project. Research data archive. Fort Collins, CO: U.S. 1945 Department of Agriculture, Forest Service, Rocky Mountain Research Station. 1946 Young, M.K.; Isaak, D.J.; Schwartz, M. [et al.]. 2018. Species occurrence data from the 1947 aquatic eDNAtlas database. Research data archive. Fort Collins, CO: U.S. Department of 1948 Agriculture, Forest Service, Rocky Mountain Research Station. 1949 https://www.fs.fed.us/rm/boise/AWAE/projects/eDNAtlas.html. (30 March 2017).

DRAFT 46

Hudec et al. Chapter 3

1950

DRAFT 47

Hudec et al. Chapter 4 48

1951 Chapter 4: Effects of Climatic Variability and Change on Forest 1952 Vegetation in Southwest Washington 1953 1954 1955 Jessica L. Hudec, Joshua S. Halofsky, Jessica E. Halofsky, Joseph A. Gates, Thomas E. DeMeo, 1956 and Douglas A. Glavich1 1957 1958 1959 Introduction 1960 1961 The Southwest Washington Adaptation Partnership (SWAP) assessment area covers a region 1962 with complex topography and volcanic geology, extending generally along Interstate 5 on the 1963 west, Interstate 90 on the north, the Cascade Range crest on the east, and the Columbia River on 1964 the south (fig. 4.1). Gifford Pinchot National Forest (GPNF) covers 457,600 ha (25 percent) of 1965 the assessment area. A small part of the southeastern portion of GPNF is excluded. The 1966 assessment area also includes Mt. Rainier National Park, a portion of Okanogan-Wenatchee 1967 National Forest, and forested lands managed by the Washington Department of Natural 1968 Resources and private timber companies. Vegetative communities in this area vary with geology, 1969 topography, climate, and disturbance regimes (fig. 4.2). 1970 Climate, specifically temperature and precipitation distribution, interacts with geology, 1971 landforms, topography and soils to create biophysical environments that favor different 1972 assemblages of plant species in time and space. These broad biophysical environments, or 1973 vegetation zones (also called “series”), are named for the dominant overstory species that would 1974 dominate the landscape without disturbance (Henderson et al. 1989). Reflecting in part a steep 1975 precipitation gradient (fig. 4.3), the Southwest Washington landscape supports several vegetation 1976 zones (fig. 4.4). The western hemlock (Tsuga heterophylla [Raf.] Sarg.) zone (Franklin and 1977 Dyrness 1988) dominates low- to mid-elevation forests with a maritime climate west of the 1978 Cascade crest. Mild, maritime climate facilitates growth of dense forests of long-lived coniferous 1979 species like western hemlock, western redcedar (Thuja plicata Donn ex D. Don), and Douglas-fir 1980 (Pseudotsuga menziesii [Mirb.] Franco). Both eastside and westside Douglas-fir zones occupy 1981 low- to mid-elevation warm forests affected by a drier, continental climate. The grand fir (Abies 1982 grandis [Douglas ex D. Don] Lindl.) zone also experiences a more continental climate and 1983 occurs at low- to mid-elevations in the Columbia River Gorge and in the easternmost portion of

1 Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650; Joshua S. Halofsky is a landscape ecologist, Washington State Department of Natural Resources, 1111 Washington St. SE, Olympia, WA 98504; Jessica E. Halofsky is a research ecologist, University of Washington, College of the Environment, School of Environmental and Forest Sciences, Box 352100, Seattle, WA 98195-2100; Joseph A. Gates is a forester and the forest vegetation program manager, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, 501 E 5th Street #404, Vancouver, WA 98661; Thomas E. DeMeo is the regional ecologist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Region, 1220 SW 3rd Avenue, Portland, OR 97204; and Douglas A. Glavich is a botanist and ecologist, U.S. Department of Agriculture, Forest Service, Air Resource Management Program, 3200 SW DRAFTJefferson Way, Corvallis, OR 97331.

Hudec et al. Chapter 4 49

1984 GPNF. The Pacific silver fir (Abies amabilis Douglas ex. J. Forbes) zone occupies mid- to high- 1985 elevation forests with cool climates, except in dry locations, where the subalpine fir (Abies 1986 lasiocarpa [Hook.] Nutt.) zone dominates. The mountain hemlock (T. mertensiana [Bong.] 1987 Carrière) and subalpine fir zones occupy high-elevation forests. 1988 Cold winters and deep snowpack in the mountain hemlock and subalpine fir zones 1989 influence lower density forests than found in other zones. Forests dominated by mountain 1990 hemlock and subalpine fir transition to subalpine meadows and parklands above continuous 1991 forest where cold winter temperatures and short growing seasons limit tree growth and forest 1992 distribution. The parkland zone is composed of a mosaic of tree islands, dwarf-shrubs, forbs, and 1993 grasses (Douglas 1972, Franklin and Dyrness 1988, Henderson 1974). The alpine zone occupies 1994 the highest elevations, above subalpine meadows and parklands and above treeline, and may 1995 include patches of sedge-turf communities, heather species, talus slopes, fellfields, and wetlands 1996 (Douglas and Bliss 1977, Edwards 1980). Elevation of continuous forest and treeline varies with 1997 latitude and aspect, reflecting differences in mean seasonal temperatures across the complex 1998 topography of the Cascades (Körner and Paulsen 2004). 1999 A large body of research exists on the potential effects of climatic variability and change 2000 on vegetation in the Pacific Northwest (Peterson et al. 2014), providing a foundation for 2001 understanding climate change vulnerabilities and for developing vegetation management and 2002 adaptation options. Effects of climate change on vegetation depend on the magnitude of changes 2003 in climate (i.e., exposure), as well as the sensitivity of species and ecological processes to those 2004 changes (Parry et al. 2007). Adaptive capacity of species, habitats, and ecological processes can 2005 reduce vulnerability. Land management agencies can facilitate adaptation through a variety of 2006 means, one of which is active forest management. 2007 In this chapter, we describe projected changes in vegetation under three potential future 2008 climate scenarios. We focus on current vegetation zones, disturbance regimes, potential future 2009 changes within vegetation zones, and management considerations for GPNF. Chapter 8 presents 2010 options for adapting forest vegetation management planning and practices to climate change. 2011 2012 2013 Potential Climate Change Effects 2014 2015 Climate change is expected to alter vegetation structure and composition, terrestrial ecosystem 2016 processes, and the delivery of important ecosystem services in future decades (Peterson et al. 2017 2014a,b). Climate influences the spatial distribution of major vegetation biomes, the abundance 2018 of species and communities within biomes, biotic interactions, and the geographic ranges of 2019 individual species. Climate also influences the rates at which terrestrial ecosystems process 2020 water, carbon, and nutrients and deliver ecosystem services like fresh water. Finally, climate 2021 influences the disturbance processes that shape vegetation structure and composition, and altered 2022 disturbance regimes under future climate projections will likely be the most important catalyst 2023 for vegetation change. Thus, climate-induced vegetation changes have important implications for 2024 wildlife habitat, biodiversity, hydrology, future disturbance regimes, and the ability of 2025 ecosystems to absorb and sequester carbon from the atmosphere. 2026 Several information sources are useful for assessing potential climate change effects on 2027 vegetation and future forest composition and structure, including long-term paleoecological 2028 records, evidence from experimental and observational studies, and simulation model projections 2029 for the future.DRAFT

Hudec et al. Chapter 4 50

2030 2031 2032 Paleoecological Records 2033 2034 Paleoecological records from the Pacific Northwest and elsewhere show that during historical 2035 warm periods, many tree species moved poleward and upward in elevation (Whitlock 1992, 2036 Whitlock and Bartlein 1997). Shifts in species distributions involved changes in species 2037 abundance rather than species extirpation; shifting distributions represent leading-edge dynamics 2038 rather than trailing-edge contraction. For example, during a warmer period in the 19th century, 2039 western hemlock became dominant in areas where Pacific silver fir and mountain hemlock were 2040 common on Mt. Rainier (Dunwiddie 1986), suggesting that western hemlock has the capacity to 2041 shift to higher elevations in a warmer climate (Zolbrod and Peterson 1999). 2042 The paleoecological record from the Pacific Northwest shows that species in stressed 2043 environments with adaptations to frequent disturbance have persisted during past periods of rapid 2044 climate change (e.g., Whitlock 1992). For the Pacific Northwest, these species include red alder 2045 (Alnus rubra Bong.), Douglas-fir, and lodgepole pine (Pinus contorta var. latifolia Engelm. ex S. 2046 Watson). Warmer and drier conditions at lower elevations in Southwest Washington would 2047 likely result in expansion of the range of Douglas-fir. Other species that may expand their ranges 2048 under these conditions include western white pine (Pinus monticola Douglas ex D. Don), Oregon 2049 white oak (Quercus garryana Douglas ex Hook.), and giant chinquapin (Chrysolepis 2050 chrysophylla [Douglas ex Hook.] Hjelmq.). Increased disturbance may also lead to range 2051 expansion of red alder. The paleoecological record for the Pacific Northwest also suggests that 2052 warming may increase the abundance of some species, including western hemlock and subalpine 2053 fir. 2054 2055 2056 Modern Records 2057 2058 Before climate-induced changes in species distribution become apparent, changes in patterns of 2059 species establishment, growth, and mortality often occur, and these preliminary changes may 2060 eventually lead to broader range shifts (Littell et al. 2008). Dendroecological (tree-ring) records 2061 show that individual tree growth and net primary productivity are sensitive to annual changes in 2062 climate in the Pacific Northwest (Ettl and Peterson 1995, Graumlich et al. 1989, Hessl and 2063 Peterson 2004, Holman and Peterson 2006, Littell et al. 2008, Nakawatase and Peterson 2006, 2064 Peterson and Peterson 2001). Effects of future climate change on both tree growth and 2065 establishment will likely differ by species, owing to varied physiologies and allocation patterns, 2066 and with elevation and topography (Ettl and Peterson 1995, Holman and Peterson 2006). 2067 At high elevations in Southwest Washington, tree growth and establishment are limited 2068 by snowpack depth and duration and growing-season length. For example, mountain hemlock 2069 growth in the Pacific Northwest is limited by spring snowpack depth and low summer 2070 temperatures (Peterson and Peterson 2001). Similarly, growth of subalpine fir in the wetter 2071 portions of its range is negatively correlated with winter precipitation and spring snowpack depth 2072 (Peterson et al. 2002). Increasing temperatures with climate change could lead to more 2073 precipitation falling as rain rather than snow, earlier snowmelt, and thus lower snowpacks and 2074 longer growing seasons (Elsner et al. 2010). Longer growing seasons could lead to increased tree 2075 establishmentDRAFT and alleviate growth-limiting factors, resulting in increased growth and

Hudec et al. Chapter 4 51

2076 productivity in some high-elevation forests. However, local site factors like soil texture and 2077 depth as well as exposure will also affect tree establishment and growth rates. 2078 Increasing temperatures, lower winter snowpack, and earlier spring snowmelt will likely 2079 result in decreased soil moisture in some areas, including on the westside of the Cascade Range 2080 (Elsner et al. 2010). In dry portions of Southwest Washington, including the rain shadow of the 2081 Olympic Range, the southeastern corner of GPNF, and portions of the Columbia River Gorge, 2082 tree growth and establishment are already limited by low summer soil moisture. Further 2083 reduction in summer soil moisture may increase drought stress to tree species currently limited 2084 by lack of summer water supply, such as Douglas-fir at low elevations (Littell et al. 2008, 2085 Restaino et al. 2016). Increased drought stress will likely result in decreased tree growth and 2086 forest productivity, particularly in currently drier forest types. 2087 Tree-ring records and modern fire history both show that years with frequent fire and 2088 broad fire extent are associated with warmer and drier spring and summer conditions in the 2089 western United States (Heyerdahl et al. 2008, Littell et al. 2009, McKenzie et al. 2004, Taylor et 2090 al. 2008). Warmer spring and summer conditions lead to relatively early snowmelt and lower 2091 summer soil and fuel moistures. Wildfire area burned in mountainous areas in western U.S. is 2092 positively related to low precipitation, drought, and high temperatures in the 20th century (Littell 2093 et al. 2009). Therefore, increased temperatures and drought occurrence in the Pacific Northwest 2094 from climate change will likely lead to increased fire frequency and extent. More frequent fires 2095 would favor fire-tolerant tree species or those that readily regenerate after fire, such as Douglas- 2096 fir and lodgepole pine, at the expense of less fire-tolerant species, such as western hemlock. 2097 Nonetheless, individual trees can withstand considerable climatic variation, and mature trees can 2098 persist for centuries. Thus, disturbance is required to facilitate change because shifts in 2099 distribution and abundance of forest species rely on disturbance events that result in high levels 2100 of mortality. 2101 Shifts in tree species distribution with climatic warming in recent decades have been 2102 documented at several locations, including western Europe (Grabherr et al. 2009) and southern 2103 California (Kelly and Goulden 2008). Consistent with paleoecological records, these shifts have 2104 generally involved movement upward in elevation or poleward. Upward elevational movement 2105 of treelines has been documented in mountainous locations across the world, including locations 2106 in Canada (Luckman and Kavanagh 2000), Sweden (Kullman 2001), Bulgaria (Meshinev et al. 2107 2000), Russia (Moiseev and Shiyatov 2003), and New Zealand (Wardle and Coleman 1992). 2108 Meta-analysis of treeline response to recent warming at 166 sites (from around the world, 2109 but mostly in North America and Europe) found that treelines at sites with more winter warming 2110 were more likely to have advanced than treelines at sites with less winter warming (Harsch et al. 2111 2009). In addition, treelines with a diffuse form, characterized by decreasing tree density with 2112 increasing altitude or latitude, were more likely to have advanced than those with an abrupt form, 2113 characterized by a continuous canopy with no decline in density right up to treeline. High 2114 elevation areas in Southwest Washington potentially affected by changing treeline dynamics 2115 include portions of Mount St. Helens, Mt. Adams, Mt. Margaret Backcountry, Mt. Rainier, Goat 2116 Rocks Wilderness, Indian Heaven Wilderness, Tatoosh Wilderness, and William O. Douglas 2117 Wilderness. Treelines in these areas are more commonly diffuse than abrupt, and climatic stress 2118 factors likely influence tree survival and growth. Furthermore, terrain is commonly steep with 2119 unstable and minimally developed soils. Thus, vegetative shifts near treeline may be highly 2120 variable and will depend on local climate and microsite conditions (e.g., soil moisture) 2121 (MalansonDRAFT et al. 2007).

Hudec et al. Chapter 4 52

2122 Drought-related tree mortality and frequency of some drought-related disturbance events 2123 have increased with recent climatic warming (e.g., Breshears et al. 2005, 2009). Moisture stress 2124 may leave trees in Southwest Washington more susceptible to attack by insects, as has been 2125 observed in other locations in western North America (e.g., Hicke et al. 2006). Drought 2126 conditions and warm, dry spring weather tend to increase tree stress factors and insect success 2127 (USDA FS and WA DNR 2016). Damage typically occurs in both the year of drought and the 2128 following year. Forest health aerial detection surveys in Washington have mapped a variety of 2129 mortality agents in Southwest Washington in recent years (USDA FS and WA DNR 2011-2016), 2130 including insects whose population numbers and success rates are positively affected by warmer 2131 temperatures and drought: mountain pine beetle (Dendroctonus ponderosae Hopkins), Douglas- 2132 fir beetle (Dendroctonus pseudotsugae Hopkins), and balsam woolly adelgid (Adelges piceae 2133 Ratzeburg). 2134 Insect outbreaks, such as that of the mountain pine beetle, have been recorded across a 2135 broad spectrum of latitudes and temperature regimes in western North America throughout 2136 history. However, the severity and distribution of some recent outbreaks differ from what can be 2137 inferred from historical records; higher temperatures associated with climate change are believed 2138 to be a significant factor in these differences (Aukema et al. 2008, Carroll et al. 2004, Logan and 2139 Powell 2001). For example, the expansive mountain pine beetle outbreak in British Columbia 2140 (1999-2015, peaking in 2004) (BC MFLNRO 2017) extended into northern areas of British 2141 Columbia and into areas east of the Rocky Mountains in Alberta, where mountain pine beetles 2142 were not successful in the past because of cold winter temperatures (Carroll et al. 2004). 2143 Lodgepole pine trees above 1,300 m elevation on Mt. Adams have experienced extensive 2144 mortality as a result of mountain pine beetle outbreaks in the past six years (USDA FS and WA 2145 DNR 2011-2016), with some lower elevation ponderosa pine (Pinus ponderosa Douglas ex P. 2146 Lawson & C. Lawson) trees also affected. Warmer temperatures and more moisture stress could 2147 result in mountain pine beetle attacks extending to other areas in Southwest Washington with a 2148 high component of lodgepole pine that are currently wetter than Mt. Adams, such as portions of 2149 Goat Rocks Wilderness and Indian Heaven Wilderness. 2150 Greater likelihood of disturbance by fire and extreme weather events in a warmer climate 2151 may influence the frequency of Douglas-fir beetle attacks, and an increase in summer drought 2152 potential may prolong outbreaks (inferred from USDA FS 2011). Douglas-fir beetle outbreaks in 2153 the western United States are typically triggered by disturbance events, such as fire or windthrow 2154 (Anderson and Palik 2011, USDA FS 2011). Beetle populations build up in weakened and fallen 2155 trees, then attack and kill live trees. Outbreaks subside within two to four years if no subsequent 2156 disturbance occurs to introduce new susceptible trees, although outbreaks may be prolonged 2157 during drought conditions (USDA FS 2011). 2158 On the westside of the Cascade Range of Oregon, windstorm events and prolonged 2159 drought were historically the most important predictors of Douglas-fir beetle outbreaks, and 2160 outbreaks were more common in areas with mature and old-growth trees (Powers et al. 1999). 2161 The number of standing live trees killed in an outbreak depend on the proportion of Douglas-fir 2162 in the area, tree density, tree age, and relative moisture stress of trees at the time of attack 2163 (USDA FS 2011). Dense, mature forests composed primarily of Douglas-fir are abundant in 2164 Southwest Washington, which may make this area susceptible to Douglas-fir beetles. 2165 Increased temperatures associated with climate change will likely lead to lower fuel 2166 moisture, longer fire seasons, and an increased frequency of years with widespread fire across 2167 the West (LittellDRAFT et al. 2010). Greater disturbance frequency and area affected, coupled with

Hudec et al. Chapter 4 53

2168 higher likelihood of summer drought, could increase the frequency, extent, and duration of 2169 Douglas-fir beetle-caused mortality in susceptible forests in Southwest Washington. 2170 Furthermore, insect outbreaks alter fuel conditions, which increase fire hazard for around five 2171 years following an outbreak (Hicke et al. 2012). 2172 Nonnative balsam woolly adelgid can infest both subalpine fir and Pacific silver fir in 2173 Southwest Washington. This insect generally does not kill trees quickly, but can cause the slow 2174 demise of infested trees. Mitchell and Buffam (2001) observed that 3-4 years of warmer than 2175 average summers resulted in increased adelgid damage in subalpine fir at higher elevations in 2176 Oregon and Washington, concluding that a long-term (decades) increase in summer temperatures 2177 would expand the range of adelgids. Currently, balsam woolly adelgid primarily affects 2178 subalpine fir and some Pacific silver fir on the south and southwest sides of Mt. Adams. A 2179 warmer climate might aggravate the effects of adelgids on subalpine fir and perhaps on Pacific 2180 silver fir, which inhabits a much larger area in Southwest Washington. 2181 Climate change may influence the incidence of tree disease in Southwest Washington, 2182 but the effects of climate change on host physiology, adaptation or maladaptation, and 2183 population genetics that affect host-pathogen interactions are poorly understood (Kliejunas et al. 2184 2009). Nevertheless, we can use existing knowledge of tree diseases in western North America to 2185 infer that climate change will result in reductions in tree health and advantageous conditions for 2186 some pathogens (Kliejunas et al. 2009). Warmer, drier summers will probably favor some root 2187 and canker diseases (Frankel et al. 2014). Armillaria root disease and cytospora canker of alder 2188 are examples of pathogens known to exist in Southwest Washington that may increase in severity 2189 under a warmer climate (Kliejunas et.al. 2009). 2190 The effects of climate change on host trees, insects, pathogens, and interactions among 2191 them will have mostly adverse consequences on forest ecosystems (Kliejunas et al. 2009). 2192 Although current insect- and pathogen-related tree mortality is relatively low in Southwest 2193 Washington, soil moisture stress is expected to increase as the climate warms, and multiple 2194 disturbance-stressor interactions could form stress complexes that influence ecosystem change 2195 (McKenzie et al. 2009). 2196 Some plant and animal species in different ecosystems across the world have begun to 2197 respond to warming over the last few decades (Parmesan 2006, Parmesan and Yohe 2003, Root 2198 et al. 2003). Most plant responses to recent warming have involved alteration of species 2199 phenologies (timing of life history stages) (Bradley et al. 1999, Parmesan 2006). For example, 2200 seasonal advances in timing of flowering have been reported for plant species in Great Britain 2201 (Fitter and Fitter 2002). In addition, the growing season has lengthened across the entire 2202 Northern Hemisphere in the last 50 years (Parmesan 2006). Shifts in timing of flowering and the 2203 abundance of insect pollinators could lead to a decline in some plant species if pollinators are 2204 absent during times of peak flowering. Specific plants and pollinators most sensitive to such 2205 shifts in timing in Southwest Washington have not yet been identified, but focal species in 2206 natural and cultural resource management (e.g., huckleberries) are of particular interest. 2207 Increased disturbance, disturbance interactions, and compounded stresses to native 2208 species may increase opportunities for establishment by nonnative plant species (Joyce et al. 2209 2008). Nonnative species can reduce native biodiversity compete with native species for 2210 resources, alter soil chemistry, influence disturbance frequency and severity, and displace native 2211 plants and wildlife that rely on them. 2212 2213 DRAFT

Hudec et al. Chapter 4 54

2214 Model Output and Projections 2215 2216 Climate-informed state-and-transition simulation models (cSTSM) (Halofsky et al. 2013, 2217 2014a,b) project potential future disturbance and changes to vegetation composition and 2218 structure under different climate scenarios and fire suppression assumptions. These models 2219 incorporate potential changes in fire regimes and vegetation type from the MC2 dynamic global 2220 vegetation model (Bachelet et al. 2001) under different climate scenarios from global climate 2221 models. Results of simulations with cSTSMs developed for the western Washington Cascades 2222 are described here. 2223 There global climate models (GCMs) were selected as inputs to MC2. The selected 2224 GCMs were chosen from the Coupled Model Intercomparison Project Phase 5 (CMIP5, 2225 http://cmip-pcmdi.llnl.gov/cmip5): HadGEM2_ES, CSIRO_MK360, and NORESM1 (HadGEM, 2226 CSIRO, and NorESM, respectively, hereafter). All GCMs were run under the representative 2227 concentration pathway (RCP) 8.5, which is characterized by relatively high future carbon dioxide 2228 concentrations in the atmosphere. We selected this subset of GCMs based on their performance 2229 for the Pacific Northwest, as evaluated in Rupp et al. (2013), and to span a range of potential 2230 future climate conditions. For the 2070–2099 period, (1) NorESM projects the least warming and 2231 the wettest future compared to the baseline 1979–2008 time period (+3.8 °C and +220 mm mean 2232 annual precipitation) (2) CSIRO projects a warmer and slightly wetter future (+4.8 °C and +58 2233 mm mean annual precipitation), and (3) HadGEM projects the hottest and driest future (+6°C 2234 and -167 mm mean annual precipitation) (see chapter 2). Temperatures increase across seasons 2235 for all three models, with higher increases in the summer. Precipitation increases are generally 2236 highest in winter, with decreases during summer, although precipitation projections are more 2237 variable than temperature in all models and are highly uncertain. 2238 The selected GCM outputs were downscaled to 30 arc-second spatial resolution using the 2239 delta method (Fowler et al. 2007). Future vegetation conditions were simulated by MC2 for 2240 2010-2100 using the downscaled GCM data as input. Detailed modeling methods are described 2241 in Halofsky et al. (2013). 2242 Output from MC2 suggests that forest productivity is likely to increase in Southwest 2243 Washington, but aboveground carbon is projected to decrease with increased wildfire (fig. 4.5). 2244 MC2 projected a 400-500 percent increase in annual area burned by wildfire depending on the 2245 GCM. MC2 was run without fire suppression because fire occurrence and effective suppression 2246 of large fires to date have been minimal in westside forests, and the success of future fire 2247 suppression efforts is uncertain. Therefore, these projections may overestimate future fire. Any 2248 potential effects from insects and pathogens were also excluded from MC2. 2249 Projections generally showed vegetation zones shifting to higher elevations (fig. 4.6). 2250 Douglas-fir and grand fir zones replaced the western hemlock zone in the warmest and driest 2251 parts of the analysis area, including areas adjacent to the Columbia River and the Puget Trough. 2252 Lower elevation portions of the Pacific silver fir zone were replaced by western hemlock, and 2253 Pacific silver fir in turn moved into portions of the mountain hemlock zone. The mountain 2254 hemlock zone moved into the subalpine parkland zone (fig. 4.6). 2255 Historically (1979–2008), the dominant vegetation zones in the study area were western 2256 hemlock (60 percent of the landscape), followed by Pacific silver fir (20 percent), and mountain 2257 hemlock (13.5 percent). On average across GCMs, western hemlock was still projected by MC2 2258 to comprise around 60 percent of the landscape for the 2070–2099 period, but Pacific silver fir 2259 was reducedDRAFT to 16 percent and mountain hemlock to 6 percent (fig. 4.7). Projections also showed

Hudec et al. Chapter 4 55

2260 that the Douglas-fir zone expanded from 2 to 15 percent of the landscape. There was 60 percent 2261 agreement in vegetation projections among the three future climate scenarios for the end of the 2262 21st century (fig. 4.8). 2263 Projections for vegetation type change and fire activity from MC2 were used as inputs to 2264 cSTSMs to examine possible future changes in forest composition and structure. We ran 2265 cSTSMs with no fire suppression as well as with a 50 percent reduction in area burned, assuming 2266 partial success of continued fire suppression. Unlike MC2, insect and pathogen transitions could 2267 affect vegetation in the cSTSMs. However, we did not alter potential effects from such 2268 disturbances under a changing climate because of limited scientific understanding of 2269 relationships between most insects and pathogens with climate. In general, fire occurrence and 2270 vegetation changes are dampened in cSTSMs compared to MC2 (fig. 4.7). Wildfire occurrence 2271 increased 200–300 percent in cSTSM projections absent fire suppression, and increased 110-210 2272 percent with fire suppression (compared to a 400-500 percent increase in MC2). 2273 There are two reasons for this difference between MC2 and cSTSM projections. First, in 2274 MC2, wildfire is deterministic for a given fire return interval, occurring whenever specific 2275 temperature and precipitation thresholds are exceeded and sufficient fuel is present. In cSTSMs, 2276 wildfire is probabilistic, occurring with a given frequency over a simulation period. Second, we 2277 allowed the possibility of a vegetation type change to occur in cSTSMs only following stand- 2278 replacing disturbances, when conditions are most conducive to plant establishment. Lowered fire 2279 frequency in cSTMs and few opportunities for vegetation type change resulted in relatively 2280 minor changes in vegetation types over the simulation period (figs. 4.7, 4.9). 2281 Relatively little change in area was projected for mountain hemlock, Pacific silver fir, 2282 and western hemlock vegetation zones through the end of the century in cSTSM simulations (fig. 2283 4.9), although the extent of each zone is likely to shift with changes in climate. In contrast, the 2284 subalpine parkland was projected to decrease by approximately half in area because of increased 2285 wildfire, warmer temperatures, and less snow. Suppressing fires greatly reduced the amount of 2286 loss in area in the subalpine parkland zone. The driest zone, Douglas-fir, increased most in area 2287 because of higher temperature and lower summer precipitation (fig. 4.9). Broad-scale vegetation 2288 changes occur gradually in this modeling approach; however, large, stand-replacing disturbance 2289 events are required to facilitate transitions, and changes in zone area will likely be abrupt 2290 following disturbance. 2291 2292 2293 Climate Change Effects on Forest Productivity 2294 2295 Site productivity measures a product (e.g., wood fiber) that can be realized at a certain site under 2296 a specified management regime. Changes in climate may generate stress on individual 2297 organisms, communities, and ecosystem processes, while resilience to climate-induced changes 2298 will influence future site productivity. Management approaches may need to be adjusted in 2299 response to changes in disturbance regimes, nutrient availability, and soil moisture. Stand-scale 2300 management parameters, including maximum stand density and maximum basal area, may shift 2301 depending on stress complexes and species responses. Standard operating procedures for tree 2302 regeneration may also need to be re-evaluated. 2303 Some studies suggest that increasing levels of free atmospheric carbon dioxide may 2304 increase productivity (e.g., Bragg et al. 2013). Free-air carbon dioxide enrichment experiments 2305 done in theDRAFT 1990s showed that increased carbon dioxide resulted in increased site productivity

Hudec et al. Chapter 4 56

2306 (DeLucia and Thomas 2000, Hymus et al. 1999, Naidu and DeLucia 1999) and improved water 2307 use efficiency in some tree species, even during periods of high moisture stress (Herrick and 2308 Thomas 1999). 2309 A study in British Columbia (Wang et al. 2006) showed lodgepole pine productivity 2310 increased as mean annual temperature rose by 5 oC, at which point productivity began to 2311 decrease. This response was attributed to resiliency thresholds of local genetic material to 2312 changes in temperature. Assuming resiliency thresholds in genetic material apply to other 2313 species, those with low genetic adaptability would likely have the lowest resiliency thresholds 2314 and therefore be most sensitive to climate change (table 4.1). 2315 2316 2317 Effects of Climate Change on Vegetation Zones 2318 2319 Because potential natural vegetation zones (the expected state of mature vegetation in the 2320 absence of human intervention or disturbance) are commonly used in vegetation management, 2321 we have included a discussion of climate change effects by potential natural vegetation zones, 2322 also noting some common indicator species within each zone. Indicator species have limited 2323 environmental distribution and are therefore representative of specific biophysical environments. 2324 Each species, and each organism within a species, has a different capacity to respond to 2325 biophysical change based on genetic characteristics. Individual organisms at a site may respond 2326 differently to the same changes, and we cannot assume species assemblages on a site will persist 2327 or that current communities will be representative of a particular vegetation zone in the future. 2328 Rather, the distribution and abundance of individual species and plant communities on the 2329 landscape are transient in space and time. Some species, such as understory plant species with 2330 relatively short life spans and lichens (box 4.1), may respond more quickly to a warmer climate 2331 than trees. 2332 2333 2334 Alpine Zone 2335 2336 The alpine zone covers approximately 2,000 ha in Southwest Washington (less than 1 percent of 2337 land) and 500 ha in GPNF (less than 1 percent of land). Vegetative communities of the alpine 2338 zone occupy the highest elevations at which vegetation is able to persist. Alpine communities 2339 occur above 2,000 m elevation and are isolated because occurrence is restricted to high mountain 2340 peaks: Mt. Adams, Mount St. Helens (fig. 4.10), Goat Rocks, and Mt. Rainier. The alpine zone is 2341 shaped by environmental conditions that limit the establishment and persistence of vegetation, 2342 including short and variable growing seasons, prolonged snowpack, extreme temperatures, high 2343 winds, and intense solar radiation (Evers et al. 1996, Franklin and Dyrness 1988). Harsh 2344 environmental conditions, dynamic topography, and poorly developed soils control vegetation 2345 patterns, growth, and reproductive processes (Major and Taylor 1988), resulting in a patchy 2346 mosaic of vegetation types and rocky substrate (Evers et al. 1996). Vegetation is characterized by 2347 a general absence of trees and thinly distributed patches of subshrubs, sedge turf, and fellfields 2348 (Douglas and Bliss 1977, Edwards 1980). Environmental and physical factors may support 2349 krummholz trees and subalpine parkland extending into the alpine zone. 2350 Disturbances in the alpine zone are common, generally local, and dominated by 2351 avalanches,DRAFT wind, rock slides, talus slippage, and stream movement. These disturbances

Hudec et al. Chapter 4 57

2352 contribute to selective pressures for vegetation establishment and persistence. Fire is uncommon 2353 in the alpine zone because of sparse fuel conditions, and natural fires are often confined to 2354 individual trees or small patches following lightning strike ignitions (Evers et al. 1996). The 2355 main interaction between fire and the alpine environment is at treeline. High-severity fire can 2356 increase alpine area by lowering treeline until conditions are favorable for conifer recovery. 2357 Alpine plants are also sensitive to recreation impacts, including trampling by hikers, rock 2358 climbers, and campers, although recreation impacts are typically limited to a small number of 2359 locations around developed recreation areas and trails (Billings 1988) (fig. 4.11). 2360 Alpine vegetation is expected to be sensitive to changes in climate because of potential 2361 for altered hydrologic regimes, possible effects on reproduction, isolation, and limited adaptive 2362 capacity (Canonne et al. 2007, Holtmeier and Broll 2005, Loarie et al. 2009, Pepin et al. 2015). 2363 Research has shown that changes in climate and hydrology affect the cover and composition of 2364 alpine plant communities (Arft et al. 1999, Cannone et al. 2007, Chapin et al. 1995, Grabherr et 2365 al. 2009, Harte and Shaw 1995, Theurillat and Guisan 2001, Walker et al. 2006, Walther et al. 2366 2002). Short growing season, poor soil conditions, and frequent disturbance hinder reproductive 2367 success of alpine species. Isolated and endemic populations have lower adaptive capacity and a 2368 higher risk of extinction than species with range continuity (Beniston 2003, Fischlin et al. 2007). 2369 Furthermore, increased disturbance and novel climatic and hydrologic conditions could favor 2370 establishment of nonnative plants. 2371 Conservation of alpine regions has scientific value because they often represent “islands” 2372 of specific components of biodiversity that are elsewhere across mountainous landscapes. 2373 Management concerns in alpine areas typically focus on human impacts from recreational 2374 activities, including hiking and camping. Disturbed alpine habitats may require active restoration 2375 for successful re-vegetation along trails and at popular recreation sites because of natural 2376 impediments to seedling establishment (chapter 8). Harsh growing conditions and isolation of 2377 alpine habitats amplify the need to incorporate locally adapted seed sources into restoration 2378 projects whenever possible. Restoration sites typically require long-term monitoring and 2379 protection during the establishment phase, because plant growth and reproduction are slow. 2380 Managers may wish to evaluate which areas are most likely to provide climate refugia in a 2381 warmer future and focus restoration efforts there. 2382 2383 2384 Parkland Zone 2385 2386 The parkland zone covers 33,400 ha in Southwest Washington (2 percent of land) and 12,600 ha 2387 of GPNF (3 percent of land). Parklands are dynamic subalpine ecotones between continuous 2388 forest and the upper limit of tree distribution (approximately 1,650–2,150 m elevation), 2389 characterized by a mosaic of vegetation types, cold winters with deep snowpack, and cool 2390 summers. Vegetation includes tree islands, ericaceous dwarf-shrubs, forbs, grasses, and 2391 wildflowers. Parklands in the analysis area are commonly associated with subalpine tree species 2392 such as mountain hemlock, subalpine fir, and whitebark pine (Diaz et al. 1997, Douglas 1970, 2393 Franklin and Dyrness 1988, Henderson 1974). Openings among tree islands provide forage and 2394 nesting areas for high-elevation wildlife such as the rosy finch (Leucosticte arctoa Pallas), white- 2395 tailed ptarmigan (Lagopus leucura Richardson), American pipit (Anthus rubescens Tunstall), and 2396 hoary marmotDRAFT (Marmota caligata Eschscholtz) (Diaz et al. 1997).

Hudec et al. Chapter 4 58

2397 Deep, persistent snowpack and wildfire drive the creation and maintenance of parklands 2398 in the Cascade Range. Snowpack persistence results in a short growing season for tree 2399 establishment, and fire maintains the mosaic of tree islands and meadows. Fires typically occur 2400 as lightning strikes in tree islands, killing individual trees or clumps of trees (LANDFIRE 2007). 2401 Burned patches can act as fuelbreaks and suppress fires that enter the parkland zone from lower 2402 elevations. Ignitions are probably common, but fires are usually less than 4 ha in size because 2403 fuels are discontinuous and a narrow range of weather and fuel conditions support active 2404 burning. Snow breakage and avalanches may be more common disturbances (LANDFIRE 2007). 2405 Patchy vegetation distribution of parklands, small fire size, and minimal research on fire history 2406 in high-elevation forests make it difficult to describe further details of the fire regime. 2407 A warmer climate may have significant effects on the parkland zone. Temperature and 2408 precipitation patterns affect tree and meadow dynamics, determine succession patterns, and 2409 control tree establishment and patch re-initiation. Increased temperature, altered hydrologic 2410 regimes, and fire exclusion have already facilitated the expansion of closed-canopy forests into 2411 parklands through increased conifer establishment over the past century (Rochefort and Peterson 2412 1986). 2413 The parkland zone may move upward in elevation as higher temperatures and changes in 2414 timing and amount of snowpack continue to favor conifer encroachment and forest-dominated 2415 systems in these high-elevation areas. Forests of mountain hemlock, lodgepole pine, and 2416 subalpine fir could overtake lower elevation parklands, and parklands might move into the alpine 2417 zone. However, space for expansion to higher elevations is limited by reduced land area 2418 available closer to mountain peaks and lack of soil development. Increased fire frequency and 2419 extent may mitigate some conifer encroachment through tree mortality, but a net loss of area in 2420 high-elevation meadows is expected. 2421 Parklands in Southwest Washington provide wildlife habitat, plant diversity, and 2422 culturally valued landscapes. Active management is restricted in most of the parkland zone, 2423 owing to wilderness designations, but prescribed fire, conifer removal, and mitigation of human 2424 impacts may be necessary to retain parkland areas in the future (chapter 8). 2425 2426 2427 Mountain Hemlock Zone 2428 2429 The mountain hemlock zone is the highest elevation forested vegetation zone, covering 242,800 2430 ha in Southwest Washington (13 percent of land) and 127,500 ha in GPNF (28 percent of land) 2431 between 1,200 and 1,800 m elevation. Much of the mountain hemlock zone is mantled by ash 2432 and pumice deposits that were eroded by glaciation and/or fluvial action and, hence, appear in a 2433 “fingered” pattern across the landscape (Diaz et al. 1997). Mountain hemlock zone also occurs 2434 on flat, high-elevation plateaus, characterized by deep snowpack and short growing seasons, with 2435 freezing temperatures possible nearly any time of year (Diaz et al. 1997). Sites often transition 2436 from cold winter conditions with heavy snow loads to dry summer conditions in a relatively 2437 short period of time. 2438 The mountain hemlock zone contains three broad biophysical environments: cold, dry; 2439 intermediate; and wet (Halofsky et al. 2013). Indicator plants in the cold, dry environment 2440 include Hitchcock’s woodrush (Luzula hitchcockii Hämet-Ahti), beargrass (Xerophyllum tenax 2441 [pursh] Nutt.), and grouse whortleberry (Vaccinium scoparium leiberg ex Coville). Indicators at 2442 intermediate,DRAFT somewhat lower (and hence warmer and more productive) elevations include

Hudec et al. Chapter 4 59

2443 queencup beadlily (Clintonia uniflora [Menzies ex Schult. & Schult. f.] Kunth) and big 2444 huckleberry (V. membranaceum Douglas ex Torr.). The wet group includes Cascade azalea 2445 (Rhododendron albiflorum Hook.) and fool’s huckleberry (Menziesia ferruginia Sm.) (Diaz et al. 2446 1997). 2447 Huckleberries and beargrass, which are valued botanical products and cultural resources 2448 for Native Americans, are commonly found in mountain hemlock forests. Native Americans 2449 historically burned these forests in zone to stimulate berry production. In the late 1800s and early 2450 1900s, European settlers started burning these forests for sheep grazing (Evers et al. 1996). 2451 Difficult access and low timber values typically preclude the mountain hemlock zone from 2452 timber harvest. 2453 Fires are relatively uncommon in the mountain hemlock zone because of long, cold 2454 winters and a short dry season, and have been relatively unaffected by fire exclusion over the 2455 past century compared to lower elevation forests that lack long-duration snowpack and have 2456 shorter natural fire return intervals. Nonetheless, feedbacks from previous fires have been shown 2457 to affect the frequency, severity, and extent of subsequent fires (Parks et al. 2014, 2015, 2016; 2458 Stevens-Rumann et al. 2014). Therefore, some degree of alteration to the natural fire regime has 2459 likely occurred in the mountain hemlock zone as a result of fire exclusion. Expansive areas of 2460 forest land in the mountain hemlock zone have burned in the Cascade Range over the past 20 2461 years, including recent large fires in GPNF: Cold Springs, 2008; Cascade Creek, 2012; and 2462 Cougar Creek, 2015 (fig. 4.12). Distinct landscape patterns have developed, including large 2463 patches of forests with high-severity fire effects. 2464 Forests of the mountain hemlock zone are classified as fire regime group V (200+ years, 2465 replacement severity) (NWCG 2015). Summer drought or an unusually dry winter may be 2466 required to create conditions favorable for fire ignition and spread. The narrow range of 2467 favorable burning conditions results in a long interval, high-severity fire regime for the mountain 2468 hemlock zone. Fires occur at intervals of 150-400 years (LANDFIRE 2016), with local ecology 2469 plot data for the Mid-Columbia Region (Mt. Hood and Gifford Pinchot National Forests and 2470 Columbia River Gorge National Scenic Area) suggesting 200-270 years (Evers et al. 1996). 2471 Evidence also exists for more frequent (50-130 year return interval), low-intensity fires in similar 2472 forest types in western Montana and central Idaho (Crane and Fischer 1986, Evers et al. 1996, 2473 Fischer and Bradley 1987). Severe fires at long intervals tend to perpetuate a dominance of 2474 lodgepole pine, whereas, low-intensity fires with sporadic and low rate of spread help maintain 2475 mountain hemlock in the overstory. 2476 The mountain hemlock zone is projected to gradually contract in area through the mid- to 2477 late-21st century (fig. 4.7, 4.8). Earlier spring snowmelt could result in a longer summer dry 2478 period, and area burned by high-severity fires would likely increase (fig. 4.5). Research suggests 2479 that mountain hemlock growth near treeline could increase (Peterson and Peterson 2001) as the 2480 energy limitation of this species is alleviated in a warmer climate (Kemp-Jennings 2017, 2481 Marcinkowski et al. 2015). Growth of mountain hemlock at lower elevations may decrease 2482 where growth is limited by low soil moisture in summer (Peterson and Peterson 2001), and the 2483 lower elevation distribution of mountain hemlock may be affected by climatic influences on 2484 disturbance regimes (Franklin et al. 1991). 2485 Less consistent snowpack in the mountain hemlock zone will have important implications 2486 for summer water availability, which could affect municipal water supplies, fish, and other 2487 aquatic organisms (chapter 3). An increase in fires in the mountain hemlock zone could improve 2488 conditionsDRAFT for non-timber forest products, including huckleberries and beargrass. Managing

Hudec et al. Chapter 4 60

2489 wildfires to promote structural diversity and variability in future fire severity and extent may be 2490 able to increase resilience of these forests to a warmer climate and associated disturbances 2491 (chapter 8). 2492 2493 2494 Subalpine Fir Zone 2495 2496 The subalpine fir zone covers 3,100 ha in Southwest Washington (less than 1 percent of land) 2497 and 2,200 ha in GPNF (less than 1 percent of land). The subalpine fir zone was not considered 2498 independently from the parkland zone in vegetation modeling for this assessment, but specific 2499 characteristics are described here. The lower extent begins around 1,500 m elevation and can 2500 extend upwards to the parkland zone. This zone is characterized by cold, moist to semi-dry 2501 climate with a short growing season and deep winter snowpack, mixed with mountain hemlock 2502 and Pacific silver fir at the wetter end of its range and typically bordered by the grand fir zone at 2503 lower elevations. Common associates include mountain hemlock, Engelmann spruce (Picea 2504 engelmannii Parry ex. Engelm.), and lodgepole pine. Whitebark pine may occur at higher 2505 elevations in the subalpine fir zone, and Douglas-fir, grand fir and western larch (Larix 2506 occidentalis Nutt.) may occur at lower elevations (Diaz et al. 1997, Franklin and Dyrness 1988). 2507 Subalpine fir can tolerate low levels of light and grow on most seedbeds, allowing seedlings to 2508 establish in the understory of mature forests. The ability to germinate and grow beneath existing 2509 trees gives subalpine fir a reproductive advantage by allowing seedlings to develop on favorable 2510 sites, sheltered from otherwise harsh environmental conditions. Although low productivity and 2511 limited access preclude timber harvest, these forests have high recreational and cultural value for 2512 hiking, camping, and hunting (Diaz et al. 1997, Franklin and Dyrness 1988). 2513 Subalpine fir trees are thin-barked and fire intolerant. Even low-intensity fires cause 2514 mortality by burning the cambium and injuring the shallow roots. Heavy branch loading that 2515 extends to the forest floor and numerous resin blisters encourage high-intensity, replacement 2516 severity fires in fir-dominated forests. Fire return intervals are typically long, similar to those 2517 mentioned for mountain hemlock zone (see above). However, proximity of subalpine fir forests 2518 to drier, lower elevation forests of the grand fir zone may increase fire frequency in some areas. 2519 Subalpine fir is typically energy limited; growth is dependent on climatic factors such as 2520 the length of the growing season, light, and temperature (Peterson et al. 2002). Warming 2521 temperatures and decreased snowpack are expected to increase subalpine fir growth near treeline; 2522 however, tree growth and seedling establishment could decrease at low-elevation sites in the 2523 subalpine fir zone in response to drought stress (Albright and Peterson 2013, Peterson et al. 2524 2002). Warmer temperatures may favor the establishment of grand fir following disturbance and 2525 result in a shift in the grand fir zone to higher elevations. 2526 Recent fires on Mt. Adams have burned a large portion of the subalpine fir zone in 2527 GPNF. Balsam woolly adelgid and fir engraver (Scolytus ventralis LeConte) had stressed or 2528 killed numerous trees prior to the fires, which contributed to high dead fuel loadings and large 2529 patches of high-severity fire effects. Any post-fire planting in burned areas should consider the 2530 species composition, genetics, and tree density that will be most favorable under future climate 2531 scenarios (chapter 8). Ponderosa pine, grand fir, and western larch might be successful at higher 2532 elevations than current ranges indicate. Genetic stock from seed zones farther south might be 2533 better adapted to future climate in Southwest Washington. Planting trees at low to moderate 2534 densities orDRAFT variable densities may encourage snowpack retention and reduce summer drought

Hudec et al. Chapter 4 61

2535 stress (Lundquist et al. 2013, Raleigh et al. 2013). Furthermore, planting of drought-tolerant 2536 bunchgrasses and other understory forbs and shrubs can stabilize soils, provide refuge for 2537 invertebrates, encourage pollinators, and assist the migration of species tolerant of a warmer 2538 climate. 2539 2540 2541 Pacific Silver Fir Zone 2542 2543 The Pacific silver fir zone covers 376,000 ha in Southwest Washington (21 percent of land) and 2544 178,300 ha in GPNF (39 percent of land). Lands classified as Pacific silver fir zone occur 2545 between 850 and 1,350 m elevation on the westside of the Cascade Range and experience a cool, 2546 moist climate (Brockway et al. 1983). This zone is bounded by the western hemlock zone below 2547 and mountain hemlock zone above. The grand fir zone or subalpine fir zone may displace the 2548 Pacific silver fir zone on the dry, east side of GPNF near the crest of the Cascade Range. 2549 Indicator shrubs and herbs within the Pacific silver zone denote specific biophysical 2550 environments. Cascade azalea occurs on cold, moist sites and does not occur outside of areas 2551 with heavy snowpack. Fool’s huckleberry and oval-leaf blueberry (Vaccinium ovalifolium Sm.) 2552 are restricted to cool sites with relatively good drainage. Devil’s club (Oplopanax horridus [Sm.] 2553 Miq.) is confined to sites with abundant water during the growing season and generally occurs at 2554 lower elevations within the zone. Salal (Gaultheria shallon Pursh) and creeping snowberry 2555 (Symphoricarpos mollis Nutt.) occur only at lower elevations on well-drained, relatively warm 2556 sites. Beargrass is widespread, with the highest cover on cold, dry sites (fig. 4.13); moist sites 2557 with deep soils are the only locations where beargrass does not occur. 2558 Moist conditions in the Pacific silver fir zone tend to abate fire hazard. Forests in the 2559 Pacific silver fir zone have been relatively unaffected by fire exclusion over the past century 2560 compared to drier forest types with shorter natural fire return intervals. However, as described 2561 for the mountain hemlock zone, feedbacks from previous fires can affect the frequency, severity, 2562 and extent of subsequent fires (Parks et al. 2014, 2015, 2016; Stevens-Rumann et al. 2014); 2563 therefore, some degree of alteration of the natural fire regime has likely occurred as a result of 2564 fire exclusion. Although fire return intervals are long, fire has historically played a role in the 2565 development of forests in this zone. Prolonged regional drought (at least 3 years), in combination 2566 with high fuel loadings and east winds, are the main drivers of fire occurrence and spread (Agee 2567 and Flewelling 1983, Evers et al. 1996). Fire occurrence also depends on lightning storm tracks 2568 and forest visitor use. 2569 Forests of the Pacific silver fir zone are classified as fire regime group V (200+ years, 2570 replacement severity) (NWCG 2015). Fire characteristics vary along temperature and moisture 2571 gradients, but fire history maps and recent wildfires suggest that most fires are either small (< 4 2572 ha) or large (> 400 ha) (Evers et al. 1996). Fire history information for Pacific silver fir forests is 2573 scarce, particularly at higher elevations. Agee (1993) specifies a fire return interval of 300-600 2574 years for higher elevation Pacific silver fir forests in the Pacific Northwest and 100-300 years for 2575 lower elevation forests. Ecology plot data for the Mid-Columbia Region indicate a fire return 2576 interval of 170-430 years (Evers et al. 1996). 2577 Fires in the Pacific silver fir zone are typically characterized by smoldering combustion 2578 and creeping rates of spread. Understory vegetation does not support fire except during 2579 prolonged drought, but recent fires in Mt. Hood National Forest and GPNF suggest that aerial 2580 fuels, suchDRAFT as mosses and lichens, can facilitate fire spread. Fires in this zone can prepare mineral

Hudec et al. Chapter 4 62

2581 seedbeds, increase scenic value, stimulate growth of botanical forest products, and improve 2582 wildlife habitat diversity by creating a mosaic of structures and age classes across the landscape 2583 (Evers et al. 1996). However, burning on cooler sites in this zone could result in reduced site 2584 fertility and productivity (Brockway et al. 1983). 2585 MC2 and cSTSMs project a decrease in extent of the Pacific silver fir zone through the 2586 year 2100 under the three future climate scenarios (figs. 4.7, 4.9). A warmer climate could favor 2587 a transition of the Pacific silver fir zone to western hemlock zone in some areas (fig. 4.7). The 2588 Pacific silver fir zone may move up in elevation, possibly replacing some area currently 2589 classified as mountain hemlock zone (fig. 4.6). Areas of Pacific silver fir zone that are currently 2590 continuous for genetic mobility and migration may become disjunct as the zone moves up in 2591 elevation toward isolated mountain peaks. An increase in area burned would result in a larger 2592 portion of the zone in the grass-forb and post-disturbance structural stage and a smaller portion 2593 in the large diameter, multi-story stage compared to projections without climate change (chapter 2594 5). Wildfire management could create more variability in horizontal and vertical structural 2595 conditions and influence future fire severity and extent. 2596 The Pacific silver fir zone provides valuable timber and non-timber forest products. 2597 Forest productivity is highly variable, but more productive, lower elevation sites that are not 2598 otherwise excluded from timber harvest are typically incorporated into the GPNF timber harvest 2599 program and vegetation management strategy. Warming temperatures and decreased snowpack 2600 are expected to increase site productivity where sufficient moisture exists. Moisture is not 2601 expected to be limiting under future climate scenarios except on the driest sites. Several notable 2602 botanical forest products occur in the Pacific silver fir zone: salal, huckleberries, and beargrass. 2603 These botanical products, along with Pacific silver fir bough sales, generate revenue for GPNF 2604 and are important traditional resources for Native Americans. Increased area burned would favor 2605 species like huckleberries and beargrass that thrive in the post-disturbance structural stage. 2606 Future climatic conditions may become more favorable to insects and pathogens that 2607 were previously constrained by cold temperatures and short growing seasons for host plants. For 2608 example, bark beetle attacks are expected to increase where winters are warmer and summers are 2609 drier (Bentz and Klepzig 2014). Warmer winters will increase the survival rate of overwintering 2610 beetles; warmer, drier summers will increase seasonal drought stress and result in favorable 2611 conditions for bark beetle growth and development. Incidence of balsam wooly adelgid on 2612 Pacific silver fir may also increase, particularly at the dry end of the zone adjacent to the 2613 subalpine fir zone. Finally, changes in soil temperature and soil moisture could increase 2614 decomposition, thereby increasing forest carbon emissions. 2615 2616 2617 Western Hemlock Zone 2618 2619 Western hemlock zone covers 1,096,200 ha in Southwest Washington (61 percent of land) and 2620 136,500 ha in GPNF (30 percent of land). Forests of the western hemlock zone occur primarily 2621 below 1,200 m elevation and in all major river drainages west of the crest of the Cascade Range. 2622 Western hemlock zone transitions to Pacific silver fir zone above 1,000 m. Western hemlock 2623 typically becomes both the dominant overstory species and the primary regenerating species in 2624 mature forests (Topik et al. 1986), but Douglas-fir is a very common early seral species. 2625 Expansive wildfires in the early 20th century, followed by extensive timber management and 2626 reforestation,DRAFT have resulted in forests dominated by various age classes of second-growth

Hudec et al. Chapter 4 63

2627 Douglas-fir. In September 1902, approximately 194,200 ha burned in and near GPNF, including 2628 the Yacolt Burn (97,100 ha), Lewis River Fire (12,000 ha), Siouxon Fire (12,000 ha), and Cispus 2629 Fire (20,200 ha) (Minore 1979). Similar large-scale fires occurred prior to the 20th century, and 2630 pollen records indicate that Douglas-fir dominance in the past millennia often coincided with 2631 charcoal peaks in the pollen profile (Brubaker 1991). 2632 Moisture and temperature are the dominant factors that affect the distribution and 2633 abundance of understory plants in the western hemlock zone (Topik et al. 1986). The wettest 2634 sites include forest wetlands with skunk cabbage (Lysichiton americanus Hultén & H. St. John), 2635 ladyfern (Athyrium filix-femina [L.] Roth), and devil’s club. Moist sites are characterized by the 2636 presence of Oregon oxalis (Oxalis oregana Nutt.) and coolwort foamflower (Tiarella trifoliata 2637 var. unifoliata [Hook.] Kurtz). Intermediate sites occupy the largest proportion of the western 2638 hemlock zone in GPNF and include relatively widespread understory plants like dwarf Oregon 2639 grape (Mahonia nervosa [Pursh] Nutt.) and vanilla leaf (Achlys triphylla). Oceanspray 2640 (Holodiscus discolor [Pursh] Maxim.), Pacific dogwood (Cornus nuttallii Audubon), and Oregon 2641 white oak (Quercus garryana Douglas ex. Hook.) are found on the driest sites. 2642 Considerable temporal and spatial variability exists in estimates of fire return intervals for 2643 the western hemlock zone (Agee 1993). Fire frequency varies along temperature and moisture 2644 gradients, and the fire record is rarely long enough or regular enough to infer any type of cyclical 2645 pattern of fire. Prolonged drought is required to dry the forest floor enough to allow fires to start 2646 and strong winds contribute to large fire spread (Evers et al. 1996). Smoldering combustion and 2647 creeping rates of spread are most common unless dry east winds stimulate increased fire 2648 behavior and accelerate fire spread. Fire history maps and recent wildfires suggest that most fires 2649 are either small (< 4 ha) or large (> 400 ha). Mid-sized fires may occur under dry conditions with 2650 light winds, although these conditions are less common, and fire suppression has likely 2651 minimized the occurrence of such fires over the past century. 2652 For classification purposes, most forests in the western hemlock zone fall into fire regime 2653 group V (200+ years, replacement severity) (NWCG 2015). Drier types may fall into fire regime 2654 group III (35-100+ years, mixed severity) or fire regime group IV (35-100+ years, replacement 2655 severity) (NWCG 2015). Fire return intervals tend to be long, and large fires are often severe 2656 where contiguous cool, moist forests are widespread or mixed with Pacific silver fir. Ecology 2657 plot data for the Mid-Columbia Region indicate average fire intervals of 50-200+ years for moist 2658 sites in the western hemlock zone (Evers et al. 1996). North of Mount St. Helens, Yamaguchi 2659 (1986) found low-severity fires occur every 40-50 years during the first 150 years of 2660 development. After the first 150 years, large, high-severity fires occur every 125-500 years. 2661 Typical fire intervals are 100-200 years for drier western hemlock forests east of the crest of the 2662 Cascade Range where they intermix with the grand fir zone (Simpson 2007). 2663 MC2 and cSTSMs models project area occupied by the western hemlock zone to be 2664 relatively static or slightly declining over the next 100 years (figs. 4.7, 4.9). Although overall 2665 area is not expected to change much, the spatial distribution on the landscape is expected to shift. 2666 A warmer climate with drier summers could favor a transition of a portion of the western 2667 hemlock zone to the westside Douglas-fir zone (fig.4.6). Western hemlock zone may move up in 2668 elevation, and displace the current lower extent of the Pacific silver fir zone (fig. 4.6). The 2669 western hemlock zone is expected to remain continuous for genetic mobility and migration as the 2670 zone moves up in elevation. 2671 The projected increase in area burned will result in a larger portion of the western 2672 hemlock zoneDRAFT in the grass-forb and post-disturbance structural stages compared to projections

Hudec et al. Chapter 4 64

2673 with no climate change (chapter 5). Proportion of area occupied by the large diameter, multi- 2674 story structural stage is projected to decrease compared to projections with no climate change, 2675 assuming wildfire occurrence and no other forest management. The current landscape in the 2676 western hemlock zone is dominated by closed canopy, mid-seral forest as a result of wildfires 2677 and timber harvest in the 20th century. Increased disturbance over time will facilitate transition of 2678 some of these mid-seral forest stands to grass-forb or possibly to large diameter, multi-story 2679 conditions, depending on fire severity. 2680 Forests in the western hemlock zone will continue to be dominated by Douglas-fir and 2681 other early-seral associates as temperature and disturbance rates increase. Shrubs could compete 2682 with tree seedlings in areas that experience multiple high-severity disturbances, particularly on 2683 drier sites where vine maple (Acer circinatum Pursh), giant chinquapin (Chrysolepis 2684 chrysophylla [Douglas ex Hook.] Hjelmq.), or snowbrush ceanothus (Ceanothus velutinus 2685 Douglas ex. Hook.) are present prior to disturbance (Simpson 2007). On the other hand, Brown 2686 et al. (2013) found no relationship between seedling abundance and shrub cover following a 2687 single burn event in the Western Oregon Cascades. Shrubs were often more abundant than tree 2688 seedlings in terms of percentage of cover, but tree seedlings were frequently taller. Irvine et al. 2689 (2009) suggest shrubs may provide a measure of heat and moisture protection for seedlings and 2690 saplings, thereby increasing establishment and survival. Such protection from harsh 2691 environmental conditions could prove advantageous to seedling survival under future climate 2692 scenarios with warmer temperatures and more frequent drought. 2693 Both productivity and decomposition rate are expected to increase where moisture is 2694 adequate. However, moisture may become limiting for tree establishment and growth on drier 2695 sites as frequency of summer drought increases. Increased moisture stress and disturbance may 2696 result in more frequent attacks by some insect species. 2697 The western hemlock zone in GPNF includes some of the highest productivity timber 2698 lands in the National Forest System. The area is characterized by large diameter, tall trees and 2699 high basal areas. Shifts of the western hemlock zone upwards in elevation and into the current 2700 range of Pacific silver fir zone are not likely to have much effect on timber production because 2701 both zones are currently actively managed for timber. Expansion of the westside Douglas-fir 2702 zone into the western hemlock zone would have negative implications for timber management 2703 because Douglas-fir zone sites tend to be less productive. Increased rates of disturbance could 2704 result in an increase in the incidence of Douglas-fir beetle attacks. Non-timber forest products in 2705 the western hemlock zone include salal and mushrooms. A reduction in area occupied by the 2706 western hemlock zone could negatively affect the collection of salal, and changes to the moisture 2707 regime would affect mushroom production. 2708 2709 2710 Eastside Douglas-fir Zone 2711 2712 No eastside Douglas-fir zone occurs in the modeled assessment area for Southwest Washington; 2713 however, approximately 900 ha (less than 1 percent of land) in the southeast portion of GPNF 2714 are classified as eastside Douglas-fir zone. Although the eastside Douglas-fir zone currently 2715 occupies only a very small portion of land in Southwest Washington, this forest type is found 2716 much more extensively to the south and east of GPNF on adjacent lands managed by the Yakama 2717 Nation, Washington Department of Natural Resources (WA-DNR), private timber companies, 2718 ColumbiaDRAFT River Gorge National Scenic Area, and Okanogan-Wenatchee National Forest and

Hudec et al. Chapter 4 65

2719 could spread westward in a warmer climate. The eastside Douglas-fir zone occupies hot, dry sites 2720 below 1,000 m elevation and is often mixed with the grand fir zone and ponderosa pine zone 2721 (Agee 1994, Barrett et al. 2010). In addition to Douglas-fir, large ponderosa pine trees may be 2722 found in the overstory, and Oregon white oak is a common early-seral associate. Basal area is 2723 low because soil moisture deficits limit plant growth during the growing season (Topik et al. 2724 1988). 2725 Several understory species are characteristic of the low soil moisture in the eastside 2726 Douglas-fir zone (Topik et al. 1988). Western fescue (Festuca occidentalis Hook.) is found on 2727 north aspects with low precipitation. Pinemat manzanita (Arctostaphylos nevadensis A. Gray) 2728 occurs on ridges and rocky sites at higher elevations in the zone and on south aspects where 2729 rocky soils lead to effectively dry conditions. Common snowberry (Symphoricarpos albus [L.] 2730 S.F. Blake) occurs on lower slope positions and protected aspects. 2731 Eastside Douglas-fir zone forests are typically classified as fire regime group I (0-35 year 2732 frequency, low and mixed severity) (NWCG 2015). Similar forests in Montana and central Idaho 2733 were found to have mean fire intervals of 10-22 years (Evers et al. 1996). Ecology plot data for 2734 the Mid-Columbia Region suggest an average fire return interval of 6-45 years (Evers et al. 2735 1996). The frequently occurring mosaic of grand fir, Douglas-fir, and ponderosa pine zones 2736 creates a combination of fire regimes, although frequent, low-intensity surface fires have 2737 historically been the most common in eastside Douglas-fir zone forests. Frequent lightning fires 2738 and Native American burning resulted in forests that were often open, park-like, and dominated 2739 by early-seral species (Agee 1994, Barrett and Arno 1982, Simpson 2007). Fire exclusion has 2740 resulted in longer periods between surface fires, higher surface and ladder fuel loadings, higher 2741 tree densities, shifts in species composition, greater potential for high-severity fire effects, and 2742 increased forest health concerns compared to pre-European settlement (Evers et al. 1996, 2743 Simpson 2007). Thinning and prescribed burning can help mitigate adverse effects of drought 2744 and severe wildfire on values at risk, including wildlife habitat for threatened and endangered 2745 species (e.g., northern spotted owl [Strix occientalis caurina Merriam]) (Jain et al. 2012, Stine et 2746 al. 2014). Recent work (Haugo et al. 2015) can help identify the specific watersheds and seral 2747 stages in need of treatment (box 4.2). 2748 Outbreaks of western spruce budworm (Choristoneura occidentalis Freeman) have 2749 occurred in the eastside Douglas-fir zone in Southwest Washington since the 1990s, particularly 2750 at the moist end of the zone where it mixes with the grand fir zone. Continued fire exclusion 2751 could favor dense multi-story conditions that support western spruce budworm populations, and 2752 lower summer precipitation would increase stress on susceptible trees. However, sites that are 2753 currently favorable to western spruce budworm may not support populations under future climate 2754 scenarios as productivity, density of live trees, and species composition change. Mountain pine 2755 beetle can be common in dense second-growth forest stands of ponderosa pine, and western pine 2756 beetle (Dendroctonus brevicomis LeConte) affects mature ponderosa pine trees, especially on the 2757 driest sites in eastside Douglas-fir zone. Increased summer moisture stress under future climate 2758 scenarios would increase susceptibility of ponderosa pine trees to these insects. Disturbance is 2759 projected to increase under future climate scenarios, and Douglas-fir beetle outbreaks are often 2760 associated with disturbance; populations spike after defoliation, wildfire, or windthrow. 2761 The eastside Douglas-fir zone was excluded from the assessment, so no quantitative 2762 results exist for this zone. However, based on projections for increased temperature, lower soil 2763 moisture, and higher area burned, eastside Douglas-fir will likely expand westward and into 2764 areas currentlyDRAFT occupied by the grand fir zone. The driest sites in the eastside Douglas-fir zone

Hudec et al. Chapter 4 66

2765 could transition to the ponderosa pine zone, although no area in Southwest Washington is 2766 currently classified as ponderosa pine. 2767 2768 2769 Westside Douglas-fir Zone 2770 2771 The westside Douglas-fir zone occupies 40,300 ha in Southwest Washington (2 percent of land), 2772 in two distinct portions of the analysis area. Most forests of the westside Douglas-fir zone occur 2773 in the northwest corner of the analysis area in the rain shadow of the Olympic Mountains and the 2774 Puget Trough, from near sea level up to 1,300 m elevation. Small amounts also occur as disjunct 2775 patches on low-elevation (below 500 m), dry sites intermixed with the western hemlock zone at 2776 the southern end of the assessment area near the Columbia River Gorge. The dominant overstory 2777 tree species is Douglas-fir. Lodgepole pine, Pacific madrone (Arbutus menziesii Pursh), western 2778 hemlock, western white pine, western redcedar, and Rocky Mountain juniper (Juniperus 2779 scopulorum Sarg.) may also occur in the northwest portion of the assessment area (Henderson et 2780 al. 1989), but are uncommon or do not exist at the southern end. Oregon white oak is a common 2781 early-seral associate throughout the zone, and shore pine (Pinus contorta var. contorta) may 2782 occur as an early seral associate in the Puget Sound area if fires are frequent. Soils have a thin 2783 organic horizon, soil texture is often coarse, and soil temperature fluctuates widely, so low soil 2784 moisture is common and can be of extended duration. Transient snowpack and sparse canopy 2785 provide little insulation from winter cold, and south-facing sites with little canopy shading 2786 experience high soil temperatures during summer. Productivity is low, and regeneration is slow 2787 and sporadic. Shrubs may compete with trees for limited resources (Henderson et al. 1989). 2788 Understory plants include dry-site shrubs like kinnickinnick (Arctostaphylos uva-ursi [L.] 2789 Spreng.) on dry, relatively high-elevation sites and oceanspray and baldhip rose (Rosa 2790 gymnocarpa Nutt.) on dry, low- to mid-elevation sites with shallow soils. Salal and sword fern 2791 (Polystichum munitum [Kaulf.] C. Presl.) are found on moderately dry, low- to mid-elevation 2792 sites with shallow, well-drained soils (Henderson et al. 1989, Topik et al. 1986). 2793 Fire is the most significant natural disturbance in the westside Douglas-fir zone, but 2794 windstorms, insects, and pathogens also affect the area. Westside Douglas-fir zone forests are 2795 classified as fire regime group III (35-100+ years, mixed severity) (NWCG 2015). Longer return 2796 intervals compared to eastside Douglas-fir forests reflects the mix of these forests with wetter 2797 forest types of the western hemlock zone (Agee 1993). Fire frequency prior to European 2798 settlement ranged from less than 10 years to 50-100 years (Chappell and Kagan 2001). Agee 2799 (1991) found a 52-76 year mean fire-return interval for Douglas-fir forests in a dry area of the 2800 western Cascades. Henderson et al. (1989) found a 138-year fire return interval for Douglas-fir 2801 zone sites in Olympic National Forest. Mixed-severity fire effects create opportunities for 2802 establishment of new cohorts of trees and increase structural complexity. Oak and madrone trees 2803 resprout after fire, but most oak-dominated sites will eventually transition to Douglas-fir- 2804 dominated forests without periodic fire (Chappell and Kagan 2001). 2805 MC2 and cSTSMs project area occupied by the westside Douglas-fir zone to increase 2806 more than any other zone (figs. 4.7, 4.9). Warmer climate with drier summers and more area 2807 burned will favor a transition of some area currently classified as western hemlock zone to 2808 Douglas-fir zone (fig. 4.6). An increase in area burned may result in a large increase in the 2809 portion of the zone in the grass/forb and post-disturbance structural stage compared to the no 2810 climate changeDRAFT scenario (chapter 5). In contrast, the proportion of area occupied by the large

Hudec et al. Chapter 4 67

2811 diameter, multi-story structural stage is projected to decrease compared to the no climate change 2812 scenario. 2813 Several unique habitats are associated with the westside Douglas-fir zone, and more 2814 frequent disturbance would have a variety of effects. Frequent fire can help maintain oak and 2815 madrone habitat and associated prairies and wetlands by killing encroaching conifers. Increased 2816 disturbance could facilitate restoration of culturally important camas (Camassia quamash [Pursh] 2817 Greene) prairies that have been degraded through fire exclusion. However, frequent fire can also 2818 remove scarce organic matter and minerals from the soil, reduce site fertility and productivity, 2819 increase competition from shrubs, and create mineral seedbeds for non-native invasive plant 2820 species. Fuels management may be required to control severity of disturbances and reduce 2821 negative fire effects to soils (chapter 8). Increased disturbance could also lead to more frequent 2822 Douglas-fir beetle attacks and an expanded distribution of nonnative plants. 2823 2824 2825 Grand Fir Zone 2826 2827 The grand fir zone covers 17,300 ha in Southwest Washington (1 percent of land) from near sea 2828 level up to 1,500 m elevation. The majority of lands classified as grand fir zone in GPNF occur 2829 in the White Salmon and Little White Salmon River drainages, which are largely excluded from 2830 the assessment area. 2831 Dry and moist subtypes of the grand fir zone represent differences in climate, disturbance 2832 regimes, and productivity. Dry subtypes tend to experience more frequent fire, whereas moist 2833 subtypes are more productive and may better support active timber management. Plant 2834 associations in the dry subtype are characterized by understory plants such as pinegrass 2835 (Calamagrostis rubescens Buckey), elk sedge (Carex geyeri Boott), and oceanspray. Indicators 2836 of moist subtypes include vine maple, vanilla leaf, thimbleberry (Rubus parviflorus Nutt.), and 2837 big huckleberry. 2838 Fire is the major disturbance agent in the grand fir zone, which is typically classified as 2839 fire regime group III (35-100+ years, mixed severity) (Evers et al. 1996, NWCG 2015), but 2840 frequency and severity of fires in this zone vary greatly. Dry subtypes in the grand fir zone may 2841 experience more frequent (9- to 25-year return interval) low-intensity fires and be classified as 2842 fire regime group I (0- to 35-year frequency, low and mixed severity) (Bork 1985, NWCG 2015). 2843 Variability in fire frequency and severity is a defining characteristic of the zone, so the range in 2844 fire return intervals is more important than the mean return interval (Halofsky et al. 2011, Stine 2845 et al. 2014). 2846 Minimal area of the grand fir zone exists in Southwest Washington, but projections 2847 indicate that the area will increase through the end of the century (figs. 4.7, 4.9). The zone is 2848 expected to expand into drier portions of the western hemlock zone, particularly in the Little 2849 White Salmon River drainage, Columbia River Gorge, and possibly Goat Rocks Wilderness. The 2850 zone could also expand upward in elevation into the subalpine fir zone on the south side of Mt. 2851 Adams. 2852 A preliminary study (DeMeo and Ringo 2010) suggests this zone will be among the most 2853 affected by increased water balance deficit during the growing season. Over time, this zone can 2854 expect more frequent summer drought and moisture stress, particularly in the dry subtype. 2855 IncreasedDRAFT drought will reduce forest productivity and favor more frequent and larger fires.

Hudec et al. Chapter 4 68

2856 Other disturbance agents in the grand fir zone include livestock grazing, nonnative 2857 invasive plants, and insects and pathogens. Grazing alters natural fire regimes by affecting fine 2858 herbaceous fuel loading and species composition. Sites that currently support grazing may shift 2859 under future climate scenarios as moisture stress and associated understory species composition 2860 changes. Nonnative plants can also affect fire regimes and alter plant-wildlife interactions. 2861 Increased rate of disturbance and greater potential for high-severity fire effects may favor the 2862 establishment of nonnative species. 2863 Insects and pathogens have been a major disturbance agent in the grand fir zone in GPNF 2864 and adjacent Yakama Nation and WA-DNR lands since the 1990s. Several outbreak cycles of 2865 western spruce budworm have caused extensive mortality and subsequent dead fuel 2866 accumulation on the south side of Mt. Adams (fig. 4.14). Continued fire exclusion could favor 2867 dense multi-story conditions that support western spruce budworm populations and additional 2868 fuel accumulation. Heavy fuel loadings with both horizontal and vertical continuity could result 2869 in high-severity fire over broad extents of land. Thinning and prescribed burning can help 2870 mitigate adverse effects of drought and severe wildfire on tree vigor and on wildlife habitat for 2871 threatened and endangered species (e.g., northern spotted owl [Strix occientalis caurina 2872 Merriam]) (Jain et al. 2012, Stine et al. 2014). Recent work (Haugo et al. 2015) can help identify 2873 specific watersheds and seral stages in need of treatment (box 4.2). 2874 2875 2876 Acknowledgments 2877 2878 The MC2 modeling described in this chapter was conducted by David Conklin and funded by the 2879 Washington Department of Natural Resources. 2880 2881 2882 Literature Cited 2883 2884 Agee, J.K. 1991. Fire history of Douglas-fir forests of the Pacific Northwest. In: Ruggiero, L.F.; 2885 Aubry, K.B.; Carey, A.B.; Mark, H., eds. Wildlife and vegetation of unmanaged Douglas-fir 2886 forests. PNW-GTR-295. Portland, OR: U.S. Department of Agriculture, Forest Service, 2887 Pacific Northwest Research Station: 25–33. 2888 Agee, J.K. 1993. Fire ecology of Pacific Northwest forests. Washington, DC: Island Press. 2889 Agee, J.K. 1994. Fire and weather disturbances in terrestrial ecosystems of the eastern Cascades. 2890 PNW-GTR-320. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 2891 Northwest Research Station. 2892 Agee, J.K.; Flewelling, R. 1983. A fire cycle model based on climate for the Olympic 2893 Mountains, Washington. Proceedings of the Seventh Conference on Fire and Forest 2894 Meteorology, April 25-28, 1983. Ft. Collins, CO: American Meteorological Society: 32–37. 2895 Albright, W.L.; Peterson, D.L. 2013. Tree growth and climate in the Pacific Northwest, North 2896 America: a broad-scale analysis of changing growth environments. Journal of Biogeography. 2897 40: 2119–2133. 2898 Anderson, P.; Palik, B. 2011. Regional examples of silvicultural adaptation strategies: Western 2899 hemlock/ Douglas-fir Forests of the Pacific Northwest. U.S. Department of Agriculture, 2900 Forest Service, Climate Change Resource Center. www.fs.usda.gov/ccrc/topics/silviculture/ 2901 pacificDRAFT-northwest. (27 October 2017).

Hudec et al. Chapter 4 69

2902 Arft, A.M.; Walker, M.D.; Gurevitch, J. [et al.]. 1999. Responses of tundra plants to 2903 experimental warming: meta-analysis of the international tundra experiment. Ecological 2904 Monographs. 69: 491–511. 2905 Aukema, B.H.; Carroll, A.L.; Zheng, Y. [et al.]. 2008. Movement of outbreak populations of 2906 mountain pine beetle: influences of spatiotemporal patterns and climate. Ecography. 31: 348– 2907 358. 2908 Bachelet, D.; Lenihan, J.M.; Daly, C. [et al.]. 2001. MC1: a dynamic vegetation model for 2909 estimating the distribution of vegetation and associated ecosystem fluxes of carbon, nutrients, 2910 and water: technical documentation version 1.0. Gen. Tech. Rep. PNW-GTR-508. Portland, 2911 OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 2912 Barrett, S. W.; Arno, S.F. 1982. Indian fires as an ecological influence in the Northern Rockies. 2913 Journal of Forestry 80: 647–651. 2914 Barrett, S.; Havlina, D.; Jones, J. [et al.]. 2010. Interagency fire regime condition class 2915 guidebook. Version 3.0. Boise, ID: U.S. Department of Agriculture, Forest Service; U.S. 2916 Department of Interior; The Nature Conservancy. https://www.frames.gov/frcc. (5 March 2917 2017). 2918 Beniston, M. 2003. Climatic change in mountain regions: a review of possible impacts. Climatic 2919 Change. 59: 5–31. 2920 Bentz, B.; Klepzig, K. 2014. Bark beetles and climate change in the United States. U.S. 2921 Department of Agriculture, Forest Service, Climate Change Resource Center. 2922 www.fs.usda.gov/ccrc/topics/insect-disturbance/bark-beetles. (25 February 2017). 2923 Billings, W. D. 1988. Alpine vegetation. In: Barbour, M.G.; Billings, W.D., eds. North 2924 American terrestrial vegetation. New York: Cambridge University Press: 391–420. 2925 Bork, J.L. 1985. Fire history in three vegetation types on the eastern side of the Oregon 2926 Cascades. Doctoral dissertation. Corvallis, OR: Oregon State University. 2927 Bradley, N.L.; Leopold, A.C.; Ross, J.; Huffaker, W. 1999. Phenological changes reflect 2928 climate change in Wisconsin. Proceedings of the National Academy of Sciences, USA. 96: 2929 9701–9704. 2930 Bragg, F.J.; Prentice, I.C.; Harrison, S.P. [et al.]. 2013. Stable isotope and modeling evidence 2931 for CO2 as a driver of glacial-interglacial vegetation shifts in southern Africa. 2932 Biogeosciences. 10: 2001–2010. 2933 Breshears, D.D.; Cobb, N.S.; Rich, P.M. [et al.]. 2005. Regional vegetation die-off in response 2934 to global-change type drought. Proceedings of the National Academy of Sciences, USA. 42: 2935 15144–15148. 2936 Breshears, D.D.; Myers, O.B.; Meyer, C.W. [et al.]. 2009. Tree die-off in response to global 2937 change-type drought: mortality insights from a decade of plant water potential measurements. 2938 Frontiers in Ecology and the Environment. 7: 185–189. 2939 British Columbia Ministry of Forests, Lands, and Natural Resource Operations [BC 2940 MFLNRO]. 2017. Responding to the impacts of the 1999-2015 mountain pine beetle 2941 outbreak. https://www2.gov.bc.ca/gov/content/industry/forestry/managing-our-forest- 2942 resources/forest-health/forest-pests/bark-beetles/mountain-pine-beetle/responding-to-the- 2943 1999-2015-outbreak. (27 Oct. 2017). 2944 Brockway, D.G.; Topik, C.; Hemstrom, M.A.; Emmingham, W.H. 1983. Plant association 2945 and management guide for the Pacific silver fir zone, Gifford Pinchot National Forest. R6- 2946 Ecol-130a-1983. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 2947 NorthwestDRAFT Region.

Hudec et al. Chapter 4 70

2948 Brown, M.J.; Kertis, J.; Huff, M.H. 2013. Natural tree regeneration and coarse woody debris 2949 dynamics after a forest fire in the western Cascade Range. Res. Pap. PNW-RP-592. Portland, 2950 OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 50 2951 p. 2952 Brubaker, L.B. 1991. Climate change and the origin of old-growth Douglas-fir forests in the 2953 Puget Sound lowland. In Ruggiero, L.F.; Aubry, K.B.; Carey, A.B.; Huff, M.H., eds. Wildlife 2954 and vegetation of unmanaged Douglas-fir forests. PNW-GTR-285. Portland, OR: U.S. 2955 Department of Agriculture, Forest Service, Pacific Northwest Research Station: 17–24. 2956 Canonne, N.; Sgorbati, S.; Guglielmin, M. 2007. Unexpected impacts of climate change on 2957 alpine vegetation. Frontiers in Ecology and the Environment. 5: 360–364. 2958 Chapin, F.S., III; Shaver, G.R.; Giblin, A.E. [et al.]. 1995. Responses of arctic tundra to 2959 experimental and observed changes in climate. Ecology. 76: 694–711. 2960 Chappell, C.B.; Kagan, J. 2001. Westside oak and dry Douglas-fir forest and woodlands. In: 2961 Johnson, D.; O’Neil, T., eds. Wildlife-habitat relationships in Oregon and Washington. 2962 Corvallis, OR: Oregon State University Press: 26–28. 2963 Crane, M.F.; Fischer, W.C. 1986. Fire ecology of the forest habitat types of central Idaho. INT- 2964 GTR-218. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain 2965 Research Station. 2966 DeLucia, E.H.; Thomas, R.B. 2000. Photosynthetic responses to CO2 enrichment of four 2967 hardwood species in a forest understory. Oecologia. 122: 11–19. 2968 DeMeo, T.E.; Ringo, C.D. 2010. A comparison of fire regime condition class assessments at 2969 multiple scales in the Pacific Northwest. In: Wade, D.D.; Robinson, M.K., eds. Proceedings 2970 of the third fire behavior and fuels conference, October 2010, Spokane, Washington, USA. 2971 Birmingham, AL: International Association of Wildland Fire, CD-ROM. 2972 DeMeo, T.E.; Swanson, F.J.; Smith, E.B. [et al.]. 2012. Applying historical fire regime 2973 concepts to forest management in the western U.S. In: Wiens, J.S.; Hayward, G.D.; Safford, 2974 H.D.; Giffen, C., eds. Historical environmental variation in conservation and natural resource 2975 management. Oxford, UK: Wiley-Blackwell: 194–204. 2976 Devine, W.; Aubry, C.; Bower, A. [et al.]. 2012. Climate change and forest trees in the Pacific 2977 Northwest: a vulnerability assessment and recommended actions for national forests. In: 2978 Sniezko, Richard A.; Man, Gary; Hipkins, Valerie; Woeste, Keith; Gwaze, David; Kliejunas, 2979 John T.; McTeague, Brianna A., tech. cords. 2017. Gene conservation of tree species— 2980 banking on the future. Proceedings of a workshop. Gen. Tech. Rep. PNW-GTR-963. 2981 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 2982 Station. p. 39. 2983 Diaz, N.M.; High, T.C.; Mellen, K.T. [et al.]. 1997. Plant association and management guide 2984 for the mountain hemlock zone: Gifford Pinchot and Mt. Hood National Forests. R6-MTH- 2985 GP-TP-08-95. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 2986 Northwest Region. 2987 Douglas, G.W. 1970. A vegetation study in the subalpine zone of the western North Cascades, 2988 Washington. Master’s thesis. Seattle, WA: University of Washington. 2989 Douglas, G.W. 1972. Subalpine plant communities of the western North Cascades, Washington. 2990 Arctic and Alpine Research. 4: 147–66. 2991 Douglas, G.W.; Bliss, L.C. 1977. Alpine and high subalpine plant communities of the North 2992 CascadesDRAFT region, Washington and British Columbia. Ecological Monographs. 47: 113–150.

Hudec et al. Chapter 4 71

2993 Dunwiddie, P.W. 1986. A 6000-year record of forest history on Mount Rainier, Washington. 2994 Ecology. 67: 58–68 2995 Edwards, O.M. 1980. The alpine vegetation of Mount Rainier National Park: structure, 2996 development and constraints. Doctoral dissertation. Seattle, WA: University of Washington. 2997 Ellis, C.J. 2007. Response of British lichens to climate change scenarios: trends and 2998 uncertainties in the projected impact for contrasting biogeographic groups. Biological 2999 Conservation 140: 217–235. 3000 Elsner, M.M.; Cuo, L.; Voisin, N. [et al.]. 2010. Implications of 21st century climate change for 3001 the hydrology of Washington State. Climatic Change. 102: 225–260. 3002 Ettl, G.J.; Peterson, D.L. 1995. Growth response of subalpine fir (Abies lasiocarpa) to climate 3003 in the Olympic Mountains, Washington, USA. Global Change Biology. 1: 213–230. 3004 Evers, L.; Hubbs, H.; Crump, R. [et al.]. 1996. Fire ecology of the Mid-Columbia region. 3005 Unpub. rep. Vancouver, WA: U.S. Department of Agriculture, Forest Service, Gifford 3006 Pinchot National Forest. 3007 Fischer, W.C.; Bradley, A.F. 1987. Fire ecology of western Montana forest habitat types. Gen. 3008 Tech. Rep. INT-GTR-223. Ogden, UT: U.S. Department of Agriculture, Forest Service, 3009 Intermountain Research Station. 3010 Fischlin, A.; Midgley, G.F.; Price, J.T. [et al.]. 2007. Ecosystems, their properties, goods, and 3011 services. In: Parry, M.L.; Canziani, O.F.; Palutikof, J.P. [et al.], eds. Climate change 2007: 3012 impacts, adaptation and vulnerability. Contribution of Working Group II to the Fourth 3013 Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, United 3014 Kingdom: Cambridge University Press: 211–272. 3015 Fitter, A.H.; Fitter, R.S.R. 2002. Rapid changes in flowering time in British plants. Science. 3016 296: 1689–1691. 3017 Fowler, H.J.; Blenkinsop, S.; Tebaldi, C. 2007. Linking climate change modeling to impacts 3018 studies: recent advances in downscaling techniques for hydrological modeling. International 3019 Journal of Climatology. 27: 1547–1578. 3020 Frankel, S.; Juzwik, J.; Koch, F. 2012. Forest tree diseases and climate change. U.S. 3021 Department of Agriculture, Forest Service, Climate Change Resource Center. 3022 www.fs.usda.gov/ccrc/topics/forest-disease. (25 February 2017). 3023 Franklin, J.F.; Dyrness, C.T. 1988. Natural vegetation of Oregon and Washington. Corvallis, 3024 OR: Oregon State University Press. 3025 Franklin, J.F.; Swanson, F.J.; Harmon, M.E. [et al.]. 1991. Effects of global climatic change 3026 on forests in northwestern North America. Northwest Environmental Journal 7:233–254. 3027 Gauslaa, Y. 2014. Rain, dew, and humid air as drivers of morphology, function, and spatial 3028 distribution in epiphytic lichens. The Lichenologist 46: 1–16. 3029 Geiser, L.H.; Neitlich, P.N. 2006. Air pollution and climate gradients in western Oregon and 3030 Washington indicated by epiphytic macrolichens. Environmental Pollution 145: 203–216. 3031 Glavich, D.A.; Geiser, L.H.; Mikulin, A.G. 2005. Rare epiphytic coastal lichen habitats, 3032 modeling and management in the Pacific Northwest. The Bryologist 108: 977–390. 3033 Grabherr, G.; Gottfried, M.; Pauli, H. 2009. Climate effects on mountain plants. Nature. 369: 3034 448. 3035 Graumlich, L.J.; Brubaker, L.B.; Grier, C.C. 1989. Long-term trends in forest net primary 3036 productivity: Cascade Mountains, Washington. Ecology. 70: 405–410. 3037 Halofsky, J.E.; Creutzburg, M.K.; Hemstrom, M.A., eds. 2014a. Integrating social, 3038 economic,DRAFT and ecological values across large landscapes. Gen. Tech. Rep. PNW-GTR-896.

Hudec et al. Chapter 4 72

3039 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 3040 Station. 3041 Halofsky, J.E., Donato, D.C.; Hibbs, D.E. [et al.]. 2011. Mixed-severity fire regimes: lessons 3042 and hypotheses from the Klamath-Siskiyou ecoregion. Ecosphere. 2: art40. 3043 Halofsky, J.S.; Halofsky, J.E.; Burcsu, T.; Hemstrom, M.A. 2014b. Dry forest resilience 3044 varies under simulated climate-management scenarios in a central Oregon, USA landscape. 3045 Ecological Applications. 24: 1908–1925. 3046 Halofsky, J.E.; Hemstrom, M.A.; Conklin, D.R. [et al.]. 2013. Assessing potential climate 3047 change effects on vegetation using a linked model approach. Ecological Modelling. 266: 3048 131–143. 3049 Hann, W.J.; Wisdom, M.J.; Rowland, M.M. 2003. Disturbance departure and fragmentation 3050 of natural systems in the Interior Columbia Basin. Res. Pap. PNW-RP-545. Portland, OR: 3051 U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 3052 Harsch, M.A.; Hulme, P.E.; McGlone, M.S.; Duncan, R.P. 2009. Are treelines advancing? A 3053 global meta-analysis of treeline response to climate warming. Ecology Letters. 12: 1040– 3054 1049. 3055 Harte, J.; Shaw, R. 1995. Shifting dominance within a montane vegetation community—results 3056 of a climate-warming experiment. Science. 267: 876–880. 3057 Haugo, R.; Zanger, C.; DeMeo, T. [et al.]. 2015. A new approach to evaluate forest structure 3058 restoration needs across Oregon and Washington, USA. Forest Ecology and Management. 3059 335: 37–50. 3060 Henderson, J.A. 1974. Composition, distribution, and succession of subalpine meadows in 3061 Mount Rainier National Park, Washington. Doctoral dissertation. Corvallis, OR: Oregon 3062 State University. 3063 Henderson, J.A.; Peter, D.H.; Lesher, R.D.; Shaw, D.C. 1989. Forested plant associations of 3064 the Olympic National Forest. R6-ECOL-TP 001-88. Portland, OR: U.S. Department of 3065 Agriculture, Forest Service, Pacific Northwest Region. 3066 Herrick, J.D.; Thomas, R.B. 1999. Effects of CO2 enrichment on the photosynthetic light 3067 response of sun and shade leaves of canopy sweetgum trees (Liquidambar styraciflua) in a 3068 forest ecosystem. Tree Physiology. 19: 779–786. 3069 Hessburg, P.F.; Agee, J.K.; Franklin, J. 2005. Dry forests and wildland fires of the inland 3070 Northwest, USA: contrasting the landscape ecology of the pre-settlement and modern eras. 3071 Forest Ecology and Management. 211: 117–139. 3072 Hessl, A.E.; Peterson, D.L. 2004. Interannual variability in aboveground tree growth in 3073 Stehekin River watershed, North Cascade Range, Washington. Northwest Science. 78: 204– 3074 213. 3075 Heyerdahl, E.K.; McKenzie, D.; Daniels, L.D. [et al.]. 2008. Climate drivers of regionally 3076 synchronous fires in the inland Northwest (1651–1900). International Journal of Wildland 3077 Fire. 17: 40–49. 3078 Hicke, J.A.; Logan, J.A.; Powell, J.; Ojima, D.S. 2006. Changing temperatures influence 3079 suitability for modeled mountain pine beetle (Dendroctonus ponderosae) outbreaks in the 3080 western United States. Journal of Geophysical Research Biogeosciences. 111: G02019. 3081 Hicke, J.A.; Johnson, M.C.; Hayes, J.L.; Preisler, H.K. 2012. Effects of bark beetle-caused 3082 tree mortalityDRAFT on wildfire. Forest Ecology and Management. 271: 81–90.

Hudec et al. Chapter 4 73

3083 Holman, M.L.; Peterson, D.L. 2006. Spatial and temporal variability in forest growth in the 3084 Olympic Mountains, Washington: sensitivity to climatic variability. Canadian Journal of 3085 Forest Research. 36: 92–104. 3086 Holtmeier, F.K.; Broll, G. 2005. Sensitivity and response of northern hemisphere altitudinal 3087 and polar treelines to environmental change at landscape and local scales. Global Ecology 3088 and Biogeography. 14: 395–410. 3089 Hymus, G.J.; Ellsworth, D.S.; Baker, N.R.; Long, S.P. 1999. Does free-air carbon dioxide 3090 enrichment affect photochemical energy use by evergreen trees in different seasons? A 3091 chlorophyll fluorescence study of mature loblolly pine. Plant Physiology 120: 1183–1191. 3092 Irvine, D.R.; Hibbs, D.E.; Shatford, J.P.A. 2009. The relative importance of biotic and abiotic 3093 controls on young conifer growth after fire in the Klamath-Siskiyou region. Northwest 3094 Science. 83: 334–347. 3095 Jain, T.B.; Battaglia, M.A.; Han, H. [et al.]. 2012. A comprehensive guide to fuel management 3096 practices for dry mixed conifer forests in the northwestern United States. Gen. Tech. Rep. 3097 RMRS-GTR-292. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 3098 Mountain Research Station. 331 p. 3099 Jones, J.; Ryan, C. 2012. Fire regime condition class mapping tool (FRCCmt), user’s guide. 3100 Boise, ID: National Interagency Fuels, Fire, and Vegetation Technology Transfer. 3101 www.niftt.gov. (5 March 2017). 3102 Keane, R.E.; Hessburg, P.F.; Landres, P.; Swanson, F.J. 2009. The use of historical range 3103 and variability (HRV) in landscape management. Forest Ecology and Management. 258: 3104 1025–1037. 3105 Kelly, A.E.; Goulden, M.L. 2008. Rapid shifts in plant distribution with recent climate change. 3106 Proceedings of the National Academy of Sciences, USA. 105: 11823–11826. 3107 Kemp-Jennings, S. 2017. Temporal and spatial variability in climate-growth response of 3108 mountain hemlock at treeline. Master’s thesis. Seattle, WA: University of Washington. 3109 Kliejunas, J.T.; Geils, B.W.; Glaeser, J.M. [et al.]. 2009. Review of literature on climate 3110 change and forest diseases of western North America. Gen. Tech. Rep. PSW-GTR-225. 3111 Albany, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research 3112 Station. 3113 Körner, C.; Paulsen, J. 2004. A world-wide study of high altitude treeline temperatures. Journal 3114 of Biogeography. 31: 713–732. 3115 Kullman, L. 2001. 20th century climate warming and tree-limit rise in the southern Scandes of 3116 Sweden. Ambio. 30: 72–80. 3117 LANDFIRE. 2007. Biophysical setting model descriptions. 3118 https://www.landfire.gov/national_veg_models_op2.php. (01 Nov. 2017). 3119 LANDFIRE. 2016. Fire regimes. https://www.landfire.gov/fireregime.php. (25 February 2017). 3120 Landres, P.B.; Morgan, P.; Swanson, F.J. 1999. Overview and use of natural variability 3121 concepts in managing ecological systems. Ecological Applications. 9: 1179–1188. 3122 Littell, J.S.; McKenzie, D.; Peterson, D.L.; Westerling, A.L., 2009. Climate and wildfire area 3123 burned in western US ecoprovinces, 1916–2003. Ecological Applications. 19: 1003–1021. 3124 Littell, J.S.; Oneil, E.E.; McKenzie, D. [et al.]. 2010. Forest ecosystems, disturbance, and 3125 climatic change in Washington State, USA. Climatic Change. 102: 129–158. 3126 Littell, J.S.; Peterson, D.L.; Tjoelker, M. 2008. Douglas-fir growth in mountain ecosystems: 3127 water DRAFTlimits tree growth from stand to region. Ecological Monographs. 78: 349–368.

Hudec et al. Chapter 4 74

3128 Loarie, S.R.; Duffy, P.B.; Hamilton, H. [et al.]. 2009. The velocity of climate change. Nature. 3129 462: 24–31. 3130 Logan, J.A.; Powell, J.A. 2001. Ghost forests, global warming, and the mountain pine beetle 3131 (Coleoptera: Scolytidae). American Entomologist. 47: 160–173. 3132 Luckman, B.; Kavanagh, T., 2000. Impact of climate fluctuations on mountain environments in 3133 the Canadian Rockies. Ambio. 29: 371–380. 3134 Lundquist, J.D.; Dickerson-Lange, S.E.; Lutz, J.A.; Cristea, N.C. 2013. Lower forest density 3135 enhances snow retention in regions with warmer winters: a global framework developed from 3136 plot-scale observations and modeling. Water Resources Research. 49: 6356–6370. 3137 Major, J.; Taylor, D.W. 1988. Alpine. In: Barbour, M.G.; Major, J., eds. Terrestrial vegetation 3138 of California. Sacramento, CA: California Native Plant Society: 601–675. 3139 Malanson, G.P., Butler, D.R., Fagre, D.B. [et al.]. 2007. Alpine treeline of western North 3140 America: linking organism-to-landscape dynamics. Physical Geography. 28: 378–396. 3141 Marcinkowski, K.; Peterson, D.L.; Ettl, G.J. 2015. Nonstationary temporal response of 3142 mountain hemlock growth to climatic variability in the North Cascade Range, Washington, 3143 USA. Canadian Journal of Forest Research. 45: 676–688. 3144 Martin, E. P. 2005. Lichen response to the environment and forest structure in the western 3145 Oregon Cascades. Doctoral dissertation. Corvallis, OR: Oregon State University. 3146 McCune, B.; Grace, J.B. 2002. Indicator species analysis. In: McCune, B.; Grace, J.B.; Urban, 3147 D.L., eds. Analysis of ecological communities. Gleneden Beach, OR: MJM Software Design: 3148 198–204. 3149 McKenzie, D.; Gedalof, Z.E.; Peterson, D.L.; Mote, P. 2004. Climatic change, wildfire, and 3150 conservation. Conservation Biology. 18: 890–902. 3151 McKenzie, D.; Peterson, D.L.; Littell, J. 2009. Global warming and stress complexes in forests 3152 of western North America. In: Bytnerowicz, A.; Arbaugh, M.J.; Riebau, A.R.; Andersen, C., 3153 eds. Wildland fires and air pollution. The Hague, Netherlands: Elsevier Publishers: 317–337. 3154 Meshinev, T.; Apostolova, I.; Koleva, E. 2000. Influence of warming on timberline rising: a 3155 case study on Pinus peuce Griseb. in Bulgaria. Phytocoenologia, 30: 431–438. 3156 Minore, D. 1979. Comparative autecological characteristics of northwestern tree species—a 3157 literature review. Gen. Tech Rep. PNW-GTR-87. Portland, OR: U.S. Department of 3158 Agriculture, Forest Service, Pacific Northwest Research Station. 3159 Mitchell, R.G.; Buffam, P.E. 2001. Patterns of long-term balsam woolly adelgid infestations 3160 and effects in Oregon and Washington. Western Journal of Applied Forestry. 16: 121–126. 3161 Moiseev, P.A.; Shiyatov, S.G. 2003. Vegetation dynamics at the treeline ecotone in the Ural 3162 highlands, Russia. In: Nagy, L.; Grabherr, G.; Körner, C.; Thompson, D.B.A., eds. Alpine 3163 biodiversity in Europe. Heidelberg, Germany: Springer: 423–435. 3164 Naidu, S.L.; DeLucia, E.H. 1999. First-year growth response of trees in an intact forest exposed 3165 to elevated CO2. Global Change Biology. 5: 609–613. 3166 Nakawatase, J.M.; Peterson, D.L. 2006. Spatial variability in forest growth-climate 3167 relationships in the Olympic Mountains, Washington. Canadian Journal of Forest Research. 3168 36: 77–91. 3169 National Wildfire Coordinating Group [NWCG]. 2015. Glossary of wildland fire 3170 terminology. http://www.nwcg.gov/term/glossary/fire-regime-groups. (25 February 2017). 3171 Parks, S.A.; Holsinger, L.M.; Miller, C.; Nelson, C.R. 2015. Wildland fire as a self-regulating 3172 mechanism: the role of previous burns and weather in limiting fire progression. Ecological 3173 ApplicationsDRAFT. 25: 1478–1492.

Hudec et al. Chapter 4 75

3174 Parks, S.A.; Miller, C.; Holsinger, L.M. [et al.]. 2016. Wildland fire limits subsequent fire 3175 occurrence. International Journal of Wildland Fire. 25: 182–190. 3176 Parks, S.A.; Miller, C.; Nelson, C.R.; Holden, Z.A. 2014. Previous fires moderate fire severity 3177 of subsequent wildland fires in two large western US wilderness areas. Ecosystems. 17: 29– 3178 42. 3179 Parmesan, C. 2006. Ecological and evolutionary responses to recent climate change. Annual 3180 Review of Ecology, Evolution, and Systematics. 37: 637–669. 3181 Parmesan, C.; Yohe, G. 2003. A globally coherent fingerprint of climate change impacts across 3182 natural systems. Nature. 421: 37–42. 3183 Parry, M.L.; Canzianai, O.F.; Palutikof, J.P. [et al.], eds. 2007. Climate change 2007: 3184 impacts, adaptation and vulnerability. A contribution of Working Group II to the Fourth 3185 Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, United 3186 Kingdom: Cambridge University Press. 3187 Pepin, N; Bradley, R.S.; Diaz, H.F. [et al.]. 2015. Elevation-dependent warming in mountain 3188 regions of the world. Nature Climate Change. 5: 424–430. 3189 Peterson, D.L.; Vose, J.M.; Patel-Weyand, T., eds. 2014a. Climate change and United States 3190 forests. Dordrecht, The Netherlands: Springer. 3191 Peterson, D.W.; Kerns, B.K.; Dodson, E.K. 2014b. Climate change effects on vegetation in the 3192 Pacific Northwest: a review and synthesis of the scientific literature and simulation model 3193 projections. Gen. Tech. Rep. PNW-GTR-900. Portland, OR: U.S. Department of Agriculture, 3194 Forest Service, Pacific Northwest Research Station. 3195 Peterson, D.W.; Peterson, D.L. 2001. Mountain hemlock growth responds to climatic 3196 variability at annual and decadal scales. Ecology. 82: 3330–3345. 3197 Peterson, D.W.; Peterson, D.L.; Ettl, G.J. 2002. Growth responses of subalpine fir (Abies 3198 lasiocarpa) to climatic variability in the Pacific Northwest. Canadian Journal of Forest 3199 Research. 32: 1503–1517. 3200 Raleigh, M.S.; Rittger, K.; Moore, C.D. [et al.]. 2013. Ground-based testing of MODIS 3201 fractional snow cover in subalpine meadows and forests of the Sierra Nevada. Remote 3202 Sensing of Environment. 128: 44–57. 3203 Restaino, C.M., Peterson, D.L.; Littell, J.S. 2016. Increased water deficit decreases Douglas-fir 3204 growth throughout western US forests. Proceedings of the National Academy of Sciences, USA. 3205 113: 9557–9562. 3206 Rochefort, R.M.; Peterson, D.L. 1996. Temporal and spatial distribution of trees in subalpine 3207 meadows of Mount Rainier National Park. Arctic and Alpine Research. 28: 52–59. 3208 Root, H.T.; Geiser, L.H.; Jovan, S.; Neitlich, P. 2015. Epiphytic macrolichen indication of air 3209 quality and climate in interior forested mountains of the Pacific Northwest. Ecological 3210 Indicators. 53: 95–105. 3211 Root, T.L.; Price, J.T.; Hall, K.R. 2003. Fingerprints of global warming on wild animals and plants. 3212 Nature. 421: 57–60. 3213 Rupp, D.E.; Abatzoglou, J.T.; Hegewisch, K.C.; Mote, P.W. 2013. Evaluation of CMIP5 20th 3214 century climate simulations for the Pacific Northwest USA. Journal of Geophysical 3215 Research-Atmospheres. 118: 10884–10906. 3216 Schmidt, K.M.; Menakis, J.P.; Hardy, C.C. [et al.]. 2002. Development of coarse-scale spatial 3217 data for wildland fire and fuel management. Gen. Tech. Rep. RMRS-GTR-87. Fort Collins, 3218 CO: U.S.DRAFT Department of Agriculture, Forest Service, Rocky Mountain Research Station.

Hudec et al. Chapter 4 76

3219 Simpson, M. 2007. Forested plant associations of the Oregon East Cascades. R6-NR-ECOL-TP- 3220 03-2007. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 3221 Region. 3222 Stevens-Rumann, C.; Prichard, S.; Morgan, P. 2014. The effects of previous wildfires on 3223 subsequent wildfire behavior and post-wildfire recovery. Northern Rockies Fire Science 3224 Network Science Review No. 1. http://nrfirescience.org/resource/12650. (12 November 3225 2017). 3226 Stine, P.; Hessburg, P.; Spies, T. [et al.]. 2014. The ecology and management of moist mixed- 3227 conifer forests in eastern Oregon and Washington: a synthesis of the relevant biophysical 3228 science and implications for future land management. Gen. Tech. Rep. PNW-GTR-897. 3229 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 3230 Station. 3231 Taylor, A.H.; Trouet, V.; Skinner, C.N. 2008. Climatic influences on fire regimes in montane 3232 forests of the southern Cascades, California, USA. International Journal of Wildland Fire. 17: 3233 60–71. 3234 Theurillat, J.P.; Guisan A. 2001. Potential impact of climate change on vegetation in the 3235 European Alps: a review. Climatic Change. 50: 77–109 3236 Topik, C. 1989. Plant association and management guide for the grand fir zone, Gifford Pinchot 3237 National Forest. R6-ECOL-TP-006-88. Portland, OR: U.S. Department of Agriculture, Forest 3238 Service, Pacific Northwest Region. 3239 Topik, C.; Halverson, N.M.; Brockway, D.G. 1986. Plant association and management guide 3240 for the western hemlock zone, Gifford Pinchot National Forest. R6-ECOL-230A-1986. 3241 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Region. 3242 Topik, C.; Halverson, N.M.; High, T. 1988. Plant association and management guide for the 3243 ponderosa pine, Douglas-fir, and grand fir zones, Mount Hood National Forest. R6-ECOL- 3244 TP-004-08. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 3245 Region. 3246 Troll, C. 1955. Der Mount Rainier und das mittlere Cascaden-gebirge. Erdkunde. 9: 264–274. 3247 United Stated Department of Agriculture, Forest Service [USDA FS]. 2002. Forest Inventory 3248 and Analysis program: forest health indicators. Broch. FS-746. Washington DC. 3249 http://www.fia.fs.fed.us/library/brochures/docs. (5 March 2017). 3250 United States Department of Agriculture, Forest Service [USDA FS]. 2008. Record of 3251 decision and final environmental impact statement and forest plan amendment 20: Gifford 3252 Pinchot National Forest and Columbia River Gorge National Scenic Area (Washington 3253 portion) site-specific invasive plant treatment project and forest plan amendment. Portland, 3254 OR: Pacific Northwest Region. 3255 United States Department of Agriculture, Forest Service [USDA FS]. 2011. Douglas-fir 3256 Beetle Management. Coeur d’Alene, ID: U.S. Department of Agriculture, Forest Service, 3257 Forest Health Protection and State Forestry Organizations. 3258 https://www.fs.usda.gov/Internet/FSE_DOCUMENTS/stelprdb5187396.pdf. (27 Oct. 2017). 3259 United States Department of Agriculture, Forest Service and Washington Department of 3260 Natural Resources [USDA FS and WA DNR]. 2011-2016. Forest health aerial survey data. 3261 Olympia, WA: Washington Department of Natural Resources. 3262 https://www.dnr.wa.gov/ForestHealth. (26 October 2017). 3263 United States Department of Agriculture, Forest Service and Washington Department of 3264 NaturalDRAFT Resources [USDA FS and WA DNR]. 2016. Forest health highlights in

Hudec et al. Chapter 4 77

3265 Washington—2016. Portland, OR: U.S. Department of Agriculture, Forest Service, Forest 3266 Health Protection and Monitoring Program. 3267 https://fortress.wa.gov/dnr/protectiongis/fmanfire/default.aspx. (26 October 2017). 3268 Walker, M.D.; Wahren, H.A.; Hollister, R.D. [et al.]. 2006. Plant community responses to 3269 experimental warming across the tundra biome. Proceedings of the National Academy of 3270 Sciences, USA. 103: 1342–1346. 3271 Walther, G.-R.; Beissner, S.; Burga, C.A. 2005. Trends in the upward shift of alpine plants. 3272 Journal of Vegetation Science. 16: 541–548 3273 Wang,T.; Hamann A.; Yanchuk, A.; O’Neill G. A.; Aitken, S.N. 2006. Use of response 3274 functions in selecting lodgepole pine populations for future climates. Global Change 3275 Biology. 12: 2404–2416. 3276 Wardle, P.; Coleman, M.C. 1992. Evidence for rising upper limits of four native New Zealand 3277 forest trees. New Zealand Journal of Botany. 30: 303–314. 3278 Whitlock, C. 1992. Vegetational and climatic history of the Pacific Northwest during the last 3279 20,000 years: implications for understanding present-day biodiversity. Northwest 3280 Environmental Journal. 8: 5–28. 3281 Whitlock, C.; Bartlein, P.J. 1997. Vegetation and climate change in northwest America during 3282 the past 125 kyr. Nature. 388: 57–61. 3283 Yamaguchi, D.K. 1986. The development of old-growth Douglas-fir forests northeast of Mount 3284 St. Helens, Washington, following an A.D. 1480 eruption. Doctoral dissertation. Seattle, 3285 WA: University of Washington. 3286 Zolbrod, A.N.; Peterson, D.L. 1999. Response of high-elevation forests in the Olympic Mountains to 3287 climatic change. Canadian Journal of Forest Research. 29: 1966–1978.

DRAFT

Hudec et al. Chapter 5 78

3288 Chapter 5: Effects of Climate Change on Special Habitats in 3289 Southwest Washington 3290 3291 3292 Jessica L. Hudec, Jessica E. Halofsky, Shiloh M. Halsey, Joshua S. Halofsky, and Daniel C. 3293 Donato1 3294 3295 3296 Introduction 3297 3298 Native plants and animals have some ability to cope with climatic variability, but the combined 3299 effects of climate change and other stressors such as habitat loss and fragmentation, can greatly 3300 affect some plant and animal species, their habitats, and biodiversity (Hannah et al. 2005, Inkley 3301 et al. 2004). Individual plant and animal species may respond either positively or negatively to 3302 both direct and indirect effects of climate change. Furthermore, related changes in habitat 3303 characteristics and quality may affect species viability. Climate change will interact with existing 3304 stressors, such as pests, pathogens, fire, and other disturbance agents, and lead to complex effects 3305 on plant and animal populations and their habitats that are difficult to predict. 3306 Climate change will have both direct and indirect effects on vegetation in Southwest 3307 Washington. Increasing temperatures and changing precipitation patterns may have direct 3308 physiological effects on some plant species (Schneider et al. 2002). Other species may be 3309 affected indirectly through climate-induced changes in phenology (Parmesan 2006); shifts in 3310 geographic ranges, species densities, and herbivore populations (Chen et al. 2011, Inkley et al. 3311 2004, Schmitz et al. 2003); and effects from other climate-related stressors such as increased fire 3312 and changing hydrologic regimes (Lawler et al. 2015). Climate change may alter plant species 3313 distributions and plant communities, which provide habitat for wildlife. In addition, there may be 3314 subtle effects of climate change on wildlife, including increases in human-wildlife interactions as 3315 people change their behavior and human populations increase in the region (e.g., more visits to 3316 higher elevation locations as summer temperatures rise) (Lawler et al. 2015). 3317 Climate change is projected to have especially significant implications for freshwater 3318 resources, affecting riparian areas, wetlands, and groundwater-dependent ecosystems. Climate 3319 change is already causing a transition from snow dominance to rain dominance at low to mid 3320 elevations in the Cascade Range, resulting in diminished snowpack and altered streamflow 3321 (Leibowitz et al. 2014, Luce et al. 2012, chapter 2). Additional effects include changes in 3322 extreme high and low flow events; alteration of groundwater recharge rates; changes in the fate 3323 and transport of nutrients, sediments, and contaminants; and temporal and spatial shifts in critical

1 Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650; Jessica E. Halofsky is a research ecologist, University of Washington, College of the Environment, School of Environmental and Forest Sciences, Box 352100, Seattle, WA 98195- 2100; Shiloh M. Halsey is the conservation science director, Cascade Forest Conservancy, 4506 SE Belmont Street, Suite 230A, Portland, OR 97215; Joshua S. Halofsky is a landscape ecologist, Washington State Department of Natural Resources, 1111 Washington St. SE, Olympia, WA 98504; and Daniel C. Donato is a forest ecologist, Washington State Department of NaturalDRAFT Resources, 1111 Washington St. SE, Olympia, WA 98504. 78

Hudec et al. Chapter 5 79

3324 ecosystem processes and functions (Johnson et al. 2012, Raymondi et al. 2013). Higher 3325 frequency and severity of droughts (Seager et al. 2007) would influence the distribution and 3326 abundance of riparian plant species and may increase susceptibility to insects attacks, as well as 3327 increase the frequency and severity of wildfires (chapter 4). 3328 Changes in species physiology, distribution, and phenology (timing of life history 3329 characteristics such as flowering) have been attributed to recent climatic warming (e.g., 3330 Parmesan 2006, Parmesan and Yohe 2003, Root et al. 2003). These responses, along with past 3331 responses evident in the paleoecological record and knowledge of species physiology and 3332 biogeography, can help in projecting how species may respond to future climate change. Using 3333 these lines of evidence, we discuss the potential direct and indirect effects of climate change on a 3334 broad range of key habitats in Southwest Washington (defined as Gifford Pinchot National 3335 Forest [GPNF] and adjacent lands to the south and west for this chapter), including (1) special 3336 habitats in forested ecosystems (late successional, early seral), (2) special habitats in non- 3337 forested ecosystems (meadows, rocky balds), and (3) special habitats associated with freshwater 3338 ecosystems (riparian, wetland, and groundwater dependent). Below, we describe potential effects 3339 of climate change on each of these special habitat groups; some species of concern that use each 3340 habitat type are highlighted (table 5.1). 3341 Management goals pertinent to GPNF include (1) maintaining a diversity of wildlife 3342 habitat and viable wildlife populations, (2) maintaining or enhancing habitat for populations of 3343 threatened, endangered, and sensitive species, (3) maintaining all riparian areas in a condition 3344 that enhances riparian-dependent resource values, and (4) managing representative areas to 3345 maintain sensitive and unique plant communities (USDA FS 1990). Knowledge of potential 3346 climate change effects on key habitats and focal species of interest informed the development of 3347 specific adaptation options designed to meet those GPNF management goals (see workshop 3348 process in chapter 1, adaptation options in chapter 8). 3349 3350 3351 Effects of Climate Change on Special Habitats in Forested Ecosystems 3352 3353 Late-Successional Forest 3354 3355 Late-successional forest ecosystems are defined by a multi-layered tree canopy with large- 3356 diameter, dominant trees often exceeding 200 years in age and abundant large-diameter snags 3357 and coarse woody debris. Late-successional forests provide nesting, denning, roosting, and 3358 foraging habitat as well as cover for many wildlife species (box 5.1). Late-successional forests 3359 are the focal habitat for key threatened species, including the northern spotted owl (Strix 3360 occidentalis caurina Merriam) and marbled murrelet (Brachyramphus marmoratus Gmelin), and 3361 other species of concern such as fisher (Martes pennanti Erxleben). 3362 The Northwest Forest Plan (USDA FS and USDI BLM 1994) designated a network of 3363 Late-Successional Reserves (LSRs) to enhance conditions of late-successional and old-growth 3364 forest ecosystems. Nine LSRs and one Managed Late-Successional Area (MLSA) occur in 3365 GPNF (fig. 5.1). These LSRs comprise 182,000 ha, or nearly one-third of the national forest, 3366 ranging in size from about 3,600 ha to 50,500 ha. GPNF provides the majority of the late- 3367 successional habitat in Southwest Washington (USDA FS 1997). 3368 Under changing climatic conditions, the most significant potential stressors to late- 3369 successionalDRAFT forests are projected increases in temperature, summer drought, insect outbreaks, 79

Hudec et al. Chapter 5 80

3370 and wildfire frequency and extent. Warmer temperature and reduced moisture availability in 3371 summer would heighten competition among trees, increasing their susceptibility to insect attack 3372 (chapter 4). Insects such as Douglas-fir beetle (Dendroctonus pseudotsugae Hopkins), currently 3373 found in isolated outbreaks, could become more prominent in late-successional forests in 3374 Southwest Washington as drought stress increases the susceptibility of Douglas-fir (Pseudotsuga 3375 menziesii [Mirb.] Franco) trees to beetle attack (Powers et al. 1999). 3376 An increase in high-severity fires and repeated fires in the same footprint, something that 3377 occurred commonly in Southwest Washington the first half of the 20th century and again during 3378 the past 20 years, could reduce the extent of late-successional forest (Davis et al. 2015) and of 3379 some physical features that contribute to structural complexity such as snags and coarse woody 3380 debris. An increase in low- to moderate-severity fires on the eastside of the Cascade Range may 3381 reduce forest density and lead to more open-canopied forests with larger residual trees (Hessburg 3382 et al. 2005) in areas like Gotchen and Peterson LSRs. Moderate-severity fires in warmer forests 3383 on the westside of the Cascade Range in Oregon have been shown to generate more age classes 3384 of Douglas-fir and offer further opportunities for shade-tolerant species to establish and develop 3385 in the mid stories (Tepley et al. 2013). Thus, an increase in low- to moderate-severity fires in 3386 westside LSRs in Southwest Washington may contribute to canopy layering and structural 3387 complexity. The overall outcome of increased fire frequency and extent may be smaller patches 3388 of complex late-successional forest mixed among patches of younger forest. 3389 In addition to influences on pests, pathogens, and wildfire activity, increases in 3390 temperature and moisture stress may also have direct impacts on the demography and structure 3391 of late-successional forests. For example, long-term research in the 500-year-old stand at the TT 3392 Munger Research Natural Area in GPNF suggests that shade-tolerant species, particularly 3393 western hemlock (Tsuga heterophylla [Raf.] Sarg.), have been experiencing higher mortality 3394 rates over the last decade than at any other time on record, including the largest size classes of 3395 trees (Donato et al., unpublished data). Because western hemlock is the most abundant tree in 3396 oldest forests that provide spotted owl and marbled murrelet habitat, substantial changes in the 3397 longevity or abundance of western hemlock could alter long-term habitat structure. 3398 We used climate-informed state-and-transition simulation models (cSTSMs) to project 3399 trends in area occupied by late-successional forests in Southwest Washington under three 3400 different future climate scenarios for the remainder of the 21st century (see chapter 4). Within 3401 each vegetation zone, we used large-diameter, multi-storied (LDM) characteristics to represent 3402 late-successional forest. LDM was defined as trees greater than 51 cm diameter at breast height, 3403 greater than 40 percent canopy cover, and multiple canopy layers. Based on cSTSM outputs, area 3404 occupied by LDM either declines or gradually increases depending on the vegetation zone, as 3405 compared to current conditions (fig. 5.2). Conversely, area in LDM increased with time in the 3406 absence of climate change for all vegetation zones. Differences in area of LDM between the 3407 “climate change” and “no climate change scenarios” are attributed to projected increases in 3408 wildfire associated with climate change. In the future, the pathway from early-seral to late- 3409 successional conditions may be less direct than it has been historically for moist forest types that 3410 would have experienced fire less frequently. 3411 The northern spotted owl (NSO), a habitat and diet specialist, is likely to be sensitive to 3412 climate change (Case 2009). Logging of low-elevation coniferous forest, especially in the 1970s 3413 and 1980s, reduced and degraded NSO habitat throughout the Pacific Northwest. Logging 3414 declined steeply following the listing of the NSO as a threatened species in 1990 and the 3415 adoption ofDRAFT the Northwest Forest Plan (USDA FS and USDI BLM 1994). In the past two 80

Hudec et al. Chapter 5 81

3416 decades, wildfires have consumed more area of NSO habitat than logging (Davis et al. 2015) 3417 (fig. 5.3), and the barred owl (Strix varia Barton) has out-competed NSO for much of the habitat 3418 that remains (Buchanan 2015, Forsman et al. 2011, Hamer 1988, Olson et al. 2005). Competition 3419 with barred owls appears to be reducing NSO populations and may be pushing NSOs into 3420 marginal habitats (Wiens et al. 2014). Climate change may add stress by affecting the quantity 3421 and quality of late-successional forest habitat and the abundance and distribution of prey species, 3422 such as the northern flying squirrel (Glaucomys sabrinus Shaw) (Gutiérrez et al. 1995). The 3423 importance of higher elevation late-successional habitat may increase if the area of lower 3424 elevation habitat decreases (Carroll 2010). 3425 The marbled murrelet, a habitat and marine diet specialist, is likely to be sensitive to 3426 climate change particularly because of its low reproductive rate. Marbled murrelets require 3427 specific habitat structures, notably large-diameter tree branches (i.e., platforms), in late- 3428 successional forests for nesting (Hamer and Nelson 1995). Nesting sites must be within 75 km of 3429 marine waters with adequate marine food sources within 5 km of the shoreline (Nelson 1997). 3430 Mineral LSR (fig. 5.1) is currently home to the only known nesting sites of marbled murrelets in 3431 GPNF. Historical logging reduced nesting habitat of the marbled murrelet; whereas, wildfires, 3432 fragmentation (which increased susceptibility to nest predation by corvids), and other incidental 3433 disturbances have contributed more to the loss of habitat in recent years. Marbled murrelets are 3434 long lived (10-15 years), so individuals may be able to survive several years of unfavorable 3435 climatic conditions. However, reproduction rates are low, so populations are slow to recover 3436 from extreme events, if they recover at all. 3437 Several species of rare lichens and bryophytes are found in close association with late- 3438 successional forests (table 5.1). Some lichens and bryophytes are specific to microclimates and 3439 may therefore be sensitive to altered temperature, moisture, and air quality. Dispersal of 3440 propagules and other living tissues by lichens and bryophytes is relatively low, which limits their 3441 ability to propagate rapidly following disturbances (box 4.1). Thus, increased disturbance by 3442 pests, pathogens, and wildfire may negatively impact rare lichens and bryophytes in late- 3443 successional forests. 3444 3445 3446 Early-Seral Pre-Forest 3447 3448 Early-seral plant communities develop after a canopy-opening disturbance and before re- 3449 establishment of a closed conifer canopy (Donato et al. 2012; Swanson et al. 2011, 2014). Early- 3450 seral “pre-forest” conditions in Southwest Washington occur as huckleberry fields, recent burns, 3451 areas affected by insects and pathogens, and intensive harvest units (large number of trees 3452 removed). Early-seral pre-forests are typified by co-dominance of shrub and herbaceous species 3453 with low conifer density, low canopy cover, and abundant snags and down wood (Swanson et al. 3454 2014). Common early-seral plant species in Southwest Washington include vine maple (Acer 3455 circinatum Pursh), huckleberry (Vaccinium spp.), red alder (Alnus rubra Bong.), quaking aspen 3456 (Populus tremuloides Michx.), Rubus spp., and various forbs and grasses. These plants provide 3457 food for black-tailed deer (Odocoileus hemionus columbianus Richardson), elk (Cervus elaphus 3458 L.), black bears (Ursus americanus Pallas), pollinators, and a variety of other wildlife species. 3459 Early-seral broadleaf tree cover serves as critical habitat for several bird species (Betts et al. 3460 2010). AbundantDRAFT snags and down wood provide habitat for birds, small mammals, herpetofauna, 81

Hudec et al. Chapter 5 82

3461 and insects through provision of cover, nesting substrate, energy, and water storage (Swanson et 3462 al. 2011, 2014). 3463 The cSTSM projections under potential future climate scenarios indicate the area 3464 occupied by early-seral pre-forest conditions (grass/forb and post-disturbance [GFP] states) is 3465 likely to increase by the mid-21st century relative to current conditions and the “no climate 3466 change scenario” for all vegetation zones analyzed (fig. 5.2), assuming no fire suppression or 3467 other forest management. Increased area burned by wildfire is responsible for the increase in 3468 GFP, and the rate of increase is higher under the climate change scenario than the no climate 3469 change scenario. Although area of early-seral pre-forest is projected to increase, reduced 3470 snowpack and summer precipitation may result in lower soil moisture, lower establishment rates, 3471 and reduced site productivity in some areas. Lower site productivity, in turn, could reduce food 3472 resources for some wildlife species. 3473 Elk and black-tailed deer are expected to be relatively insensitive to climate change, 3474 because they are forage and habitat generalists and can easily disperse to find favorable habitat. 3475 Warmer and drier conditions may favor the expansion of elk and deer associated with increased 3476 area burned by wildfire, creating more young stands and edge habitat (McKelvey and Buotte 3477 2018, Thomas et al. 1979). In addition, longer periods of mild winter weather may reduce forage 3478 stress during winter months. However, wildlife biologists at GPNF have indicated that a variety 3479 of stressors are currently affecting GPNF elk herds, including insufficient high-quality forage to 3480 support elk populations. Reduced site productivity and quality of forage could contribute to 3481 current stresses on elk. 3482 Black bears are expected to be moderately sensitive to climate change. They are forage 3483 and habitat generalists, although seasonality affects availability of important food resources and 3484 species physiology (Case 2009). Shifts in timing of fruit production in early-seral pre-forests and 3485 other calorically important foods (e.g., salmon) could cause bears to increase movements to find 3486 alternative food sources. Increased bear mobility could lead to more conflicts with humans, 3487 which historically has resulted in higher mortality for bears. Climate change projections indicate 3488 that the winter hibernation period for bears could be disrupted by periods of unseasonably warm, 3489 wet weather. Black bears can remain active during the winter, but metabolically active bears 3490 require more food resources, which are typically scarcer during the winter (Nelson et al. 1983). 3491 Although increased rates of disturbance support the expansion of early-seral pre-forests, 3492 more favorable conditions for tree establishment and growth at high elevations may affect the 3493 longevity and distribution of early-seral habitat, including huckleberry fields (fig. 5.4). 3494 Huckleberry fields support a variety of wildlife species and are a traditional cultural food source, 3495 valuable botanical product, and recreational attraction (chapter 7). 3496 3497 3498 Bigleaf Maple 3499 3500 Bigleaf maple (Acer macrophyllum Pursh) habitat occurs across Southwest Washington and is 3501 particularly abundant in the Cowlitz Valley Ranger District on GPNF. In these habitats, bigleaf 3502 maple is the dominant tree species; bracken fern (Pteridium aquilinum [L.] Kuhn), western 3503 swordfern (Polystichum munitum [Kaulf.] C. Presl) and salal (Gaultheria shallon Pursh) occupy 3504 the understory. Bigleaf maple forests provide nesting and foraging habitat for some woodpecker 3505 species andDRAFT for Neotropical migratory birds. Abundant shade, hardwood logs, and leaf litter 82

Hudec et al. Chapter 5 83

3506 create favorable habitat for salamanders and mollusks, including the Puget Oregonian snail 3507 (Cryptomastix devia Gould). 3508 Limited information exists on the sensitivity of the Puget Oregonian snail to climate 3509 change. This species is found in cool, moist forests at low to moderate elevations, especially 3510 under large woody debris and leaf litter in the bigleaf maple-western swordfern plant association. 3511 Shade provides refugia from moderate fluctuations in temperature and moisture. Increased 3512 disturbance associated with climate change may lead to increased abundance of sprouting 3513 deciduous hardwoods, including bigleaf maple, although disturbance is generally considered 3514 undesirable for preservation of Puget Oregonian habitat (Burke 1999). Snails have low mobility 3515 and are therefore particularly susceptible to direct mortality from fire (Applegarth 1995, Ray and 3516 Bergey 2015). Disturbance can also result in a reduced surface litter, understory vegetation, 3517 woody debris, fungal associations, and shade, thus reducing food availability, increasing 3518 competition, increasing risk of desiccation, and increasing predation (Applegarth 1995, Burke 3519 1994, Ray and Bergey 2015). Reduced snowpack, longer dry seasons, and increased incidence of 3520 summer drought associated with climate change could also affect habitat moisture requirements. 3521 Thus, climate change may increase the abundance of bigleaf maple trees, but specific habitat 3522 requirements for the Puget Oregonian snail may be compromised as a result of increased 3523 disturbance and altered moisture conditions, possibly restricting their range to riparian areas. 3524 Since 2011, dieback of branches and whole crowns of bigleaf maple have been observed 3525 at many locations in Southwest Washington. To date, the cause of “bigleaf maple decline” has 3526 not been identified, and it is unknown how climate change might affect it. 3527 3528 3529 Quaking Aspen 3530 3531 Upland quaking aspen habitat is found primarily on the east side of GPNF along the crest of the 3532 Cascade Range. Aspen is the characteristic and dominant tree in this habitat; scattered ponderosa 3533 pine, Douglas-fir, and grand fir (Abies grandis [Douglas ex D. Don] Lindl.) may also be present. 3534 Tall shrubs, such as beaked hazel (Corylus cornuta Marshall), vine maple, and oceanspray 3535 (Holodiscus discolor [Pursh] Maxim.) are common, and pinegrass (Calamagrostis rubescens 3536 Buckley) often dominates the surface vegetation in the absence of shrubs. Aspen provides habitat 3537 for a variety of bird species, and forage and cover for ungulates. 3538 Fire plays an important role in maintenance of aspen habitat, killing mature tree stems 3539 and inducing root sprouting, which makes aspen a competitive post-fire colonizer (Howard 3540 1996). Fire exclusion can negatively impact shade-intolerant aspen; conifers eventually dominate 3541 seral aspen stands. However, the Cold Springs Fire (2008), Cascade Creek Fire (2012), and 3542 Cougar Creek Fire (2015) on the south side of Mt. Adams have provided opportunities for aspen 3543 to re-develop in this area (fig. 5.5). Where aspen regeneration is abundant, ungulate browsing 3544 and domestic livestock grazing can significantly retard growth and recruitment of stems into the 3545 overstory (Shepperd et al. 2006). 3546 Aspen habitat exhibits sensitivity to increasing temperatures, decreased moisture 3547 availability, and altered fire regimes, so its abundance and extent may decrease with climate 3548 change (Frey et al. 2004, Marchetti et al. 2011, chapter 4). Recent declines in aspen health in 3549 some locations in the western United States may be partially explained by high temperatures and 3550 low soil moisture over the last several decades (Rehfeldt et al. 2009). Moisture will likely 3551 continue toDRAFT be a limiting factor under future climate scenarios with additional warming and more 83

Hudec et al. Chapter 5 84

3552 frequent and more extreme seasonal drought predicted. The ability of aspen to re-sprout 3553 following fire moderates its sensitivity to climate change; however, high-severity fire and 3554 multiple reburns could eliminate aspen in drier locations. Aspen dispersal ability is limited in 3555 some locations, because specific conditions are required for successful seeding and germination 3556 (Jones and DeByle 1985). Aspen are likely to continue to occur in areas expected to remain 3557 relatively moist, such as lacustrine wetlands, meadows and wetlands fed by groundwater, and 3558 meadows in deep yet wide valleys. 3559 3560 3561 Subalpine Parkland 3562 3563 Subalpine parklands occur just below treeline, typically between 1,700 and 2,200 m elevation in 3564 Southwest Washington, and primarily in wilderness areas in GPNF. This habitat appears as a 3565 mosaic of small patches of trees, treeless openings, and scattered trees (fig. 5.6). Climate in 3566 subalpine habitats is characterized by cool summers and cold winters with deep snowpack 3567 (chapter 4). 3568 Whitebark pine (Pinus albicaulis [Engelm.] Rydb.) is an ecologically important species 3569 that occurs in subalpine parklands and is a candidate species for listing under the Endangered 3570 Species Act because populations have been decreasing in abundance over the past century. 3571 Whitebark pine colonizes high-elevation sites after major disturbances. Once established, it 3572 provides a protected environment for other plants to repopulate these areas with difficult growing 3573 conditions (Tomback et al. 2001). Whitebark pine provides structural complexity, nesting 3574 habitat, shelter, and an energy-rich food source for various wildlife species, especially Clark’s 3575 nutcracker (Nucifraga columbiana A. Wilson) (Tomback and Kendall 2001). Presence of 3576 whitebark pine in the subalpine parkland effectively reduces snow melt, avalanche potential, and 3577 soil erosion (Farnes 1990). Fire exclusion has reduced whitebark pine habitat, because whitebark 3578 pine depends on fire to create bare soil and full sun for germination and establishment (fig. 5.7). 3579 White pine blister rust (Cronartium ribicola A. Dietr.), an exotic fungal pathogen, has caused 3580 extensive mortality and weakening of whitebark pine populations for several decades (Geils et al. 3581 2010, McDonald and Hoff 2001). In recent decades, mountain pine beetle outbreaks have also 3582 killed cone-bearing whitebark pine trees in some locations (Arno 1986, Waring and Six 2005). 3583 Climate change is predicted to adversely affect whitebark pine and some subalpine 3584 parkland habitats. Whitebark pine is a long-lived species that may be able to persist through 3585 considerable climatic variability, although particular requirements for successful regeneration 3586 result in relatively slow reproduction and adaptability of whitebark pine. In addition, competition 3587 from more shade-tolerant conifers may increase as the climate warms and conditions at higher 3588 elevations become more hospitable for other species. However, increased wildfire may support 3589 the persistence of whitebark pine by creating more open habitat in the subalpine zone (Loehman 3590 et al. 2011). Keane et al. (2012) provide a range-wide restoration strategy for whitebark pine 3591 across multiple scales. 3592 Clark’s nutcrackers (Nucifraga columbiana A. Wilson) are instrumental in whitebark 3593 pine regeneration after large fires, because the birds disperse seeds more effectively than wind 3594 disperses seeds of competing species (Tomback et al. 1990). Clark’s nutcrackers also tend to 3595 cache seeds below and above the current distribution of whitebark pine, thereby facilitating 3596 adaptation to altered climate (Tomback 2005). Decreased nutcracker populations at some 3597 locations DRAFTin western Washington may have been caused by lack of food where whitebark pine 84

Hudec et al. Chapter 5 85

3598 mortality has increased (McKinney and Tomback 2007, Ray et al. 2017), and further losses of 3599 whitebark pine in a warmer climate would be expected to negatively affect the birds as well. 3600 3601 3602 Giant Chinquapin 3603 3604 Giant chinquapin (Chrysolepis chrysophylla var. chrysophylla [Douglas ex Hook.] Hjelmq.) is 3605 common in northern California and Oregon but occurs as small, disjunct populations in 3606 Washington, making it a sensitive species in Washington. These disjunct populations are the 3607 northernmost occurrences of the species, possibly relicts from a historically more widespread 3608 distribution or the result of rare dispersal events (Kruckeberg 1980). Giant chinquapin occurs in 3609 both shrub and tree forms, typically on dry, open sites on slopes and ridges where it 3610 provides cover and nuts for many species of birds, small mammals, and insects (Meyer 2012). 3611 Giant chinquapin is the only known host plant for the golden hairstreak butterfly (Habrodais 3612 grunus Broisduval), a candidate species for listing in Washington. The only confirmed existence 3613 of golden hairstreak butterflies in Washington is in a single grove of giant chinquapin trees in 3614 southern Skamania County (Pyle 1989). 3615 Primary threats to giant chinquapin are competition from overtopping conifers, harvest, 3616 and conversion of forest land to other uses (Meyer 2012, Shoal 2009). Giant chinquapin can 3617 persist in some shaded environments, although it achieves better growth (Kruckeberg 1980), 3618 greater abundance (Brockman 1958), and increased sexual reproduction (Keeler-Wolf 1988) in 3619 open conditions (Meyer 2012). Disturbance may be required for giant chinquapin to remain a 3620 substantial component of any landscape, because suppression from overtopping conifers 3621 accounts for a large portion of mortality (McKee 1990). 3622 Species that are drought tolerant and fire resistant, including golden chinquapin, may 3623 become more competitive in a warmer climate with more wildfire, particularly low- to mixed- 3624 severity fire. However, high-intensity fire in dense conifer canopies or abundant coarse woody 3625 debris could result in giant chinquapin mortality. Disjunct giant chinquapin populations may be 3626 genetically distinct from the contiguous distribution; therefore, loss of populations in 3627 Washington could result in the potential loss of unique genotypes (Aubry et al. 2011). 3628 Butterflies, including the golden hairstreak, exhibit both direct (e.g., activity and 3629 emergence are influenced by temperature) and indirect sensitivity to climate, primarily due to 3630 habitat specialization. The golden hairstreak spends much of its adult life, and all of its egg, 3631 larval, and pupal life stages, in golden chinquapins, although little else is known about the habitat 3632 requirements of this butterfly species (WDFW 2015). Thus, the degree of golden hairstreak 3633 sensitivity to climate change is difficult to determine. 3634 3635 3636 Oregon White Oak 3637 3638 Oregon white oak (Quercus garryana Douglas ex Hook.) and associated vegetation comprise 3639 distinct forest and woodland habitats in Southwest Washington, specifically on dry sites 3640 throughout the Columbia River Gorge and in the White Salmon watershed on the south side of 3641 GPNF, and in native prairies of the Cowlitz and Chehalis Valleys on the north side of GPNF. 3642 Low soil moisture plays a role in maintaining varying degrees of open canopies in these Oregon 3643 white oakDRAFT habitats, and soils are typically shallow, rocky, or very deep and excessively drained 85

Hudec et al. Chapter 5 86

3644 (Chappell et al. 2001). Oregon white oak and Douglas-fir typically dominate the canopy in 3645 densely forested oak habitat, and ponderosa pine occurs on some sites. Douglas-fir tends to 3646 dominate the regeneration layer. 3647 Many moths, butterflies, gall wasps, and spiders are found exclusively in association with 3648 Oregon white oak (Larsen and Morgan 1998). Acorns, oak leaves, fungi, understory shrubs and 3649 forbs, and insects provide food for a variety of wildlife species. In return, animal dissemination 3650 of acorns may be important in Oregon white oak dispersal (Chappell et al. 2001). Oregon white 3651 oak habitats provide important roosting, nesting, and feeding habitat for Neotropical migrant 3652 birds, other birds (e.g., wild turkey [Meleagris gallopavo L.]), and mammals, including the 3653 western gray squirrel (Sciurus griseus Ord) (a threatened species in Washington State). These 3654 squirrels frequently nest in either large conifers or oaks greater than 40 cm diameter. Current 3655 distribution of western gray squirrels in Washington includes Oregon white oak woodlands and 3656 conifer forests of Klickitat and southern Yakima counties, southeast of GPNF (WDFW 2013). 3657 Oregon white oak habitat in Washington is declining in extent and condition (Chappell et 3658 al. 2001, Larsen and Morgan 1998). Historically, Oregon white oak habitats were maintained by 3659 frequent natural fires and Native American burning (Agee 1990, Chappell et al. 2001, Kertis 3660 1986, Taylor and Boss 1975, Thilenius 1968). Frequent, low-intensity fires limit conifer 3661 establishment, reduce competition from grasses, stimulate oak sprouting, and reduce fuel loads 3662 (Agee 1993). Moderate-severity fires create openings, provide opportunities for establishment of 3663 new cohorts of trees, and increase structural complexity (Chappell et al. 2001). Fire exclusion 3664 over the past century has contributed to habitat loss by promoting conifer dominance and 3665 accumulation of coarse woody debris (Agee 1993). In the absence of fire, most Oregon white 3666 oak-dominated forested habitat will eventually convert to Douglas-fir forests (Holms 1990, 3667 Sugihara and Reed 1987). Other threats to Oregon white oak habitat include degradation from 3668 human development, high-severity fires, logging, and grazing (Chappell et al. 2001, Larsen and 3669 Morgan 1998). Competition from nonnative plant species may limit Oregon white oak seedling 3670 success in unburned areas (Holms 1990). 3671 Increased drought and disturbance may facilitate sprouting of deciduous hardwoods 3672 including Oregon white oak (Halofsky et al. 2011). More frequent wildfires could also reduce 3673 herbaceous biomass and favor Oregon white oak reproduction (Holms 1990). However, fire 3674 exclusion, logging, grazing, and nonnative species have altered conditions in Oregon white oak 3675 habitats, making responses to fire and other disturbances more difficult to project. Fire exclusion 3676 and logging typically result in increased fuel loadings and horizontal and vertical continuity of 3677 fuels, creating the potential for high-intensity fire and crown fire. Moderate to heavy grazing 3678 alters fire behavior, reduces species richness of understory vegetation, compacts soils, and 3679 increases soil moisture, which may promote conifer growth and encroachment (Larsen and 3680 Morgan 1998). Any form of ground disturbance facilitates establishment by nonnative plants, 3681 which can further alter fire regimes, compete with Oregon white oak seedlings, and displace 3682 native understory plants in oak habitats. 3683 Small population numbers and low genetic diversity of western gray squirrels in 3684 Washington make the species particularly vulnerable to climate-induced changes in Oregon 3685 white oak habitat. Increased frequency of high-severity fire may reduce the number of large- 3686 diameter trees available as nesting sites. Finally, climate-induced habitat loss may interact with 3687 competition from nonnative squirrels and wild turkeys to negatively impact western gray 3688 squirrels. 3689 DRAFT 86

Hudec et al. Chapter 5 87

3690 3691 Effects of Climate Change on Special Habitats in Non-Forested Ecosystems 3692 3693 Meadows (Wet and Dry) 3694 3695 Wet, moist, and dry meadow habitats occur across Southwest Washington. In GPNF, meadows 3696 add diversity within an otherwise forested landscape (fig. 5.8). Topography influences meadow 3697 locations, and elevation influences types of vegetation that occur in the meadows, as it relates to 3698 growing season length, climate, soil development, and glacial history. Wet meadows are most 3699 common on GPNF and are particularly prominent in alpine and subalpine vegetation zones. Wet 3700 meadows (fig. 5.9) are saturated with water for much, if not all, of the growing season. Moist 3701 meadows may be flooded soon after snowmelt but may not stay saturated as the water table 3702 lowers. Dry meadows may experience intermittent flooding but are well drained and have a 3703 deeper water table than wet or moist meadows. 3704 Meadows filter sediment from runoff; provide breeding habitat for invertebrates, which 3705 serve as a food source for many birds, amphibians, and reptiles; and provide habitat structure for 3706 birds and small mammals, which are a prey base for raptors and other carnivores (USDI NPS 3707 2016). Several bird species nest at meadow edges, and mammals such as Cascade golden- 3708 mantled ground squirrels (Spermophilus saturatus Rhoads), black-tailed deer, and elk are 3709 common, modifying meadow habitat through burrow construction and foraging. Threats to 3710 meadows include altered hydrologic regimes, trampling, grazing, nonnative plants, and altered 3711 fire regimes (Jakubowski 2015, USDI NPS 2016). 3712 The habitat of the mardon skipper butterfly (Polites mardon W.H. Edwards), a 3713 Washington State endangered (federal candidate) species, exists in a total of 39 documented 3714 moist-dry upland meadows in GPNF (Jakubowski 2015). These upland meadows also support a 3715 wide assortment of other butterflies including skippers, checkerspots, fritillaries, sulphurs, blues, 3716 and swallowtails, as well as populations of rare plants, including several species of grapeferns 3717 (e.g., northern grapefern [Botrychium pinnatum H. St. John]). 3718 Seasonally wet meadows or meadow-like environments provide habitat for pale blue- 3719 eyed grass (Sisyrinchium sarmentosum Suksd. ex Greene), a Washington State threatened 3720 (federally sensitive) species, for which only 24 occurrences have been documented worldwide 3721 (Ruchty 2011) (box 5.2). All populations are in Oregon and Washington, and 9 of the 24 are in 3722 GPNF. 3723 Wet meadow habitat will likely decrease with warming climate because of projected 3724 changes in hydrology, including more precipitation falling as rain, decreased snowpack, and 3725 earlier spring snowmelt (chapter 2). These hydrologic changes effectively lengthen the dry 3726 season and increase the potential for summer drought by reducing the amount and longevity of 3727 groundwater recharge in wet and moist meadows. Furthermore, changes in hydrologic regimes 3728 may reduce opportunities for water storage, influence channel development in wet meadows, 3729 induce shifts in species assemblages, and increase erosion (Dwire and Mellmann-Brown 2017, 3730 WDFW 2015). 3731 Lower snowpack and a longer growing season in alpine and subalpine wet meadows 3732 would encourage tree establishment on meadow perimeters. Conifer encroachment associated 3733 with warmer temperatures likely represents the greatest stressor for alpine and subalpine 3734 meadows (WDFW 2015). Subalpine conifers have been shown to move into alpine meadows in 3735 the PacificDRAFT Northwest during periods of warmer climate and less snow (Rochefort and Peterson 87

Hudec et al. Chapter 5 88

3736 1996, Woodward et al. 1995). Meadow plant species may be able to expand to areas that were 3737 previously covered by ice or bare ground. However, substrates may not be favorable to plant 3738 establishment and the process of soil formation would likely take many centuries. Loss of high- 3739 elevation meadows would reduce habitat for wolverine (Gulo gulo L.), Cascade red fox (Vulpes 3740 vulpes cascadensis Merriam), American pika (Ochotona princeps Richardson), and other species 3741 associated with alpine and subalpine meadows. Earlier runoff and increased summer drought 3742 may also favor tree establishment in lower elevation wet meadows by lowering the water table. 3743 Climate change favors dry meadows, which are generally well-adapted to warm and dry 3744 conditions and periodic soil drought. Increased fire frequency and drought limitations on tree 3745 species distribution may increase area and quality of dry meadow habitat (Littell et al. 2010). 3746 However, dry meadows may be sensitive to altered fire regimes, particularly increased fire 3747 frequency or severity. Frequent high-severity fire could limit native species regeneration or 3748 increase establishment of nonnative species. Thus increased area and quality of dry meadows 3749 could positively influence species that depend on dry meadow habitat, including mardon skipper, 3750 but climate change could also lead to changes in plant species composition, increased nonnative 3751 species, and reduced habitat quality for certain species (Halofsky et al. 2011). 3752 Climate sensitivity of mardon skipper is influenced by temperature, precipitation, and fire 3753 (WDFW 2015). Temperature influences mardon skipper foraging behavior, adult life span, and 3754 larval development. Warming temperature may affect phenological timing between mardon 3755 skippers and plants that provide nectar and host-specific reproductive stages. Precipitation affects 3756 adult behavior, and extreme precipitation can cause mortality by preventing foraging or by 3757 drowning larvae. Moist conditions can also contribute to fungal development. Fire helps 3758 maintain open meadow habitat used by the mardon skipper, although this species is not very 3759 mobile, so fire can cause direct mortality of all life stages. Increasing fire frequency may expand 3760 overall habitat area available for mardon skipper, but fire in current habitat areas could 3761 contribute to local extirpation. 3762 Pale blue-eyed grass is a regionally endemic species with a narrow ecological amplitude 3763 and, therefore, may have limited ability to respond to climate change. Altered growing season, 3764 altered hydrology, and competition from nonnative species may stress pale blue-eyed grass or 3765 challenge its ability to adapt to new environmental conditions (Ruchty 2011). However, 3766 increased disturbance by fire may reduce competition from woody plants for light and water in 3767 existing pale blue-eyed grass habitat, and fire in forested stands adjacent to existing habitat could 3768 result in habitat expansion. 3769 3770 3771 Rock Outcrops 3772 3773 Rock outcrops in Southwest Washington include boulder fields, lava flows, rocky summits, 3774 cliffs, and talus slopes. Rock outcrops tend to be too steep or rocky to support trees, although 3775 patchy vegetation may occur. Seepage and pooling may provide moisture for mosses, lichens, 3776 and wetland vegetation. Overall vegetative diversity reflects the variability in soil depth and 3777 moisture content. The uniqueness of rock outcrop habitats can result in distinct communities 3778 compared to adjacent habitats and seasonal variability within a specific habitat area. High levels 3779 of diversity within and among habitats increase overall species richness of rock outcrops. 3780 A warmer climate may affect plant and animal species composition on rock outcrops. For 3781 example, DRAFTmountain goats (Oreamnos americanus Blainville) (fig. 5.10) require high, cool, rocky 88

Hudec et al. Chapter 5 89

3782 terrain for survival, with summer temperatures indirectly influencing survival of goats during the 3783 following winter (White et al. 2011). Cool summers prolong the emergence of herbaceous plants 3784 at the edge of slowly receding snowlines, thus providing food later in the summer. Plants in early 3785 growth stages also promote higher weight gains in animals. Therefore, higher summer 3786 temperatures could negatively affect mountain goat populations by reducing the amount of late- 3787 season forage. 3788 Rock outcrops also provide cover and hibernacula for reptiles, amphibians, and small 3789 mammals. Higher temperature and moisture fluctuations can negatively impact reptile and 3790 amphibian populations. Reptiles are particularly sensitive to changes in microclimates (Harvey 3791 and Weatherhead 2010), and some amphibian species need sites for burrowing in wet areas to 3792 keep their skin moist (Marks 2006). 3793 3794 3795 Rocky Balds 3796 3797 Presence of rocky balds in Southwest Washington is primarily controlled by soil conditions. 3798 Rocky balds typically occur as small openings on slopes within forested landscapes where soils 3799 are too shallow and dry to support trees. Sites can be seasonally moist or wet but tend to become 3800 extremely dry late in the growing season. Rocky balds cover a relatively small area but provide 3801 habitat for many plant species that do not occur elsewhere in western Washington (Chappell 3802 2006). 3803 Disturbances (e.g., wildfire, wildlife foraging behavior, and recreation) affect rocky bald 3804 habitats. Fires can help maintain or expand the size of balds by killing small trees. Deer, elk, and 3805 bears may cause surface disturbance on balds by grazing and digging for bulbs (Chappell 2006). 3806 High recreational use can reduce vegetation cover, cause surface erosion, and increase nonnative 3807 species. 3808 A warmer climate with more drought, combined with unfavorable soils, may limit conifer 3809 establishment and growth on rocky balds. Balds with scattered mature trees may be more 3810 susceptible to additional tree establishment because of microclimate amelioration adjacent to 3811 mature trees (Chappell 2006). Increased fire frequency and extent could expand the area of balds 3812 by killing small young trees on the margins, although increased disturbance and drought could 3813 also increase establishment of nonnative species and negatively affect species richness. 3814 3815 3816 Alpine 3817 3818 Alpine habitat occurs above 2,100 m elevation on major peaks and ridges in Southwest 3819 Washington, most notably Mount St. Helens, Mt. Adams, and Goat Rocks. Alpine habitats are 3820 partially vegetated and rugged with steep, rocky ridges, snowfields, and glaciers. Alpine habitats 3821 are subject to high winds, extreme temperature fluctuations, and intense ultraviolet radiation. 3822 Alpine vegetation occurs in patches and includes krummholz (fig. 5.11), sedge turfs, lichens, 3823 grasses, flowering herbs, and mosses. Alpine vegetation is controlled by factors such as snow 3824 cover and retention, wind desiccation, and a short growing season (Beniston 2003, Burrows 3825 1990, Körner 1998). Many alpine communities are endemic to a particular place because 3826 species’ rangeDRAFT continuity among peaks is lacking. Isolation among populations may limit genetic 89

Hudec et al. Chapter 5 90

3827 diversity and adaptive capacity of endemic alpine plant and animal communities to climate 3828 change. 3829 Many animal species that live at high elevations in the Pacific Northwest, such as 3830 American pika, hoary marmot (Marmota caligata Eschscholtz), Cascade red fox (Washington 3831 State candidate species), and wolverine (Washington State and federal candidate species) are a 3832 conservation concern because of their evolutionary histories and high sensitivity to climate 3833 change (Atkins 2012). Most pikas in the contiguous United States inhabit high-elevation 3834 ecosystems, but some survive at lower elevations, including populations in the Columbia River 3835 Gorge. Pikas cannot tolerate high temperatures, have low reproductive rates, have limited 3836 dispersal ability, and require rock fields in close proximity to meadows for habitat (Jeffress et al. 3837 2013). 3838 Hoary marmots typically live above treeline where rock slides occur next to moist 3839 meadows. Long winters without vegetation available for food dictate much of their life history, 3840 including an eight-month hibernation period each year. Steep gradients in snow cover and soil 3841 moisture in alpine habitats affect the productivity and distribution of preferred forage plant 3842 species. Marmots are sensitive to daily and seasonal fluctuations in temperature and 3843 precipitation, snow cover duration, forage availability, and predation (Leventhal 2007). Marmots 3844 have little competition for food, although finding preferred forage species may require them to 3845 travel across areas with minimal cover to protect them from predators. Predation, reproduction, 3846 dispersal, and winter starvation all affect hoary marmots. 3847 Cascade red foxes are adapted to cold climates, with habitat restricted to alpine and 3848 subalpine ecosystems with meadows. Volcanoes of the Cascade Range appear to act as islands of 3849 habitat for small, isolated Cascade red fox populations (Atkins 2012). Sensitivity of Cascade red 3850 fox to climate change is driven by dependence on these colder, high-elevation habitats and 3851 potential changes in prey abundance (WDFW 2015). 3852 Wolverines are habitat specialists and utilize very snowy habitats, which limits 3853 competition from other carnivores (Inman et al. 2012). Wolverines exhibit sensitivity to high 3854 temperature and reduced snowpack. They require persistent spring snow cover for denning, prey 3855 caching, reducing thermal stress, and reducing competition from other carnivores and scavengers 3856 (Copeland et al. 2010, Inman et al. 2012, WDFW 2015). 3857 In a warmer climate, alpine habitat may gradually migrate to higher elevations in some 3858 locations. However, changes in climate may occur at a rate that exceeds migration capacity for 3859 some species, depending on individual longevities, survival rates, and competition (Beniston 3860 2003). Climate change may affect soil availability and moisture, precipitation amount falling as 3861 snow versus rain, timing and rate of snowmelt, extent of glacial forefields, extent of permanent 3862 snowfields, disturbance regimes, seed dispersal, and germination and survival of alpine plants 3863 (Malanson et al. 2007). Although glacial recession may increase the extent of forefields, 3864 geomorphic disturbance at this interface (landslides, debris flows) may also increase erosion. 3865 Rising temperatures and shifting precipitation patterns are expected to result in reduced 3866 snowpack, thus reducing winter thermal insulation for pikas in some locations. Conversely, 3867 rising temperatures could make alpine winters less harsh and reduce the amount of snowpack 3868 needed to provide adequate seasonal insulation. Pikas that inhabit isolated mountaintops are 3869 expected to be vulnerable to climate change. Pikas at lower elevations may also be vulnerable 3870 where current habitat nears the edge of their physiological tolerance and where primary 3871 productivityDRAFT is expected to decline (Friggens et al. 2018). Overall, pikas are expected to be 90

Hudec et al. Chapter 5 91

3872 resilient to a warmer climate if sufficient habitat is available for thermal moderation and forage 3873 (Millar and Westfall 2010). 3874 Hoary marmots may be more vulnerable to a warmer climate than pikas. As snowpack 3875 decreases in the future, burrows used for hibernation may have less insulation, which could 3876 affect hoary marmot physiology and time of emergence. Potential phenological mismatches with 3877 growth of preferred vegetation used for food, decreased duration of food abundance, introduction 3878 of nonnative species, loss of habitat and habitat connectivity as a result of forest succession, and 3879 potential for increased competition or predation as alpine habitats become more favorable to 3880 other species could negatively impact hoary marmot populations (Johnston et al. 2012). Hoary 3881 marmot populations may also gradually shift to higher elevations if suitable habitat is available. 3882 Warming temperatures and declines in snowpack could affect Cascade red foxes and 3883 wolverines by decreasing habitat patch size, quality, and connectivity and by facilitating 3884 movement of less snow-adapted competitors (e.g., coyote [Canis latrans Say] and nonnative red 3885 foxes [Vulpes vulpes L.]) (Atkins 2012). Wolverines may also experience reduced success of 3886 prey caching, limited den sites or thermal refugia, and increased dispersal costs (WDFW 2015). 3887 3888 3889 Effects of Climate Change on Special Habitats in Riparian, Wetland, and 3890 Groundwater-Dependent Ecosystems 3891 3892 Riparian 3893 3894 Riparian areas are diverse, complex, and dynamic terrestrial habitats (Gregory et al. 1991, 3895 Naiman et al. 1993, Naiman and Décamps 1997). Riparian areas can be defined as “three- 3896 dimensional zones of direct physical and biotic interactions between terrestrial and aquatic 3897 ecosystems, with boundaries extending outward to the limits of flooding and upward into the 3898 canopy of streamside vegetation” (Gregory et al. 1991). These areas provide critical habitat to a 3899 diverse array of species, including Neotropical birds, ducks, amphibians, and rare botanical 3900 species (e.g., cold-water corydalis [Corydalis aquae-gelidae M. Peck and Wilson], a Washington 3901 State sensitive species). Riparian areas also influence water quality throughout the stream 3902 system. Riparian vegetation serves a variety of ecological functions for streams, including 3903 provision of shade, stream bank stability, and input of organic matter and large wood. 3904 Various terms and numerous administrative definitions and have been developed to assist 3905 in managing riparian areas (USDA FS 2012). Riparian areas, wetlands, and intermittent streams 3906 are included within USDA Forest Service Riparian Habitat Conservation Areas, which specify 3907 minimum buffer widths from each side of the channel or stream edge: intermittent streams (15 3908 m), wetlands and non-fish-bearing perennial streams (46 m), and fish-bearing streams (91 m). 3909 Active management within these buffers must comply with a number of riparian management 3910 objectives designed to improve habitat conditions for fish species that have been federally listed 3911 as threatened or endangered under the Endangered Species Act (USDA FS 1995). 3912 Riparian vegetation depends on the presence of flowing water. Streamflow can vary 3913 considerably with season, physical features of the watershed, and water source; some stream 3914 segments flow perennially, whereas others have only intermittent flow. Streamflow volume can 3915 drive seasonal changes in water table elevation of the adjacent riparian area (Jencso et al. 2011), 3916 contributingDRAFT to fluvial processes and the formation of geologic surfaces that are essential for the 91

Hudec et al. Chapter 5 92

3917 establishment, development, and persistence of different riparian plant species and communities 3918 (Dwire and Mellmann-Brown 2017, Naiman et al. 2005). 3919 Climate change may affect the seasonality, amount, and type of precipitation, and timing 3920 and rate of snowmelt (Luce et al. 2012, 2013; Safeeq et al. 2013), which would impact snowpack 3921 volumes (Hamlet et al. 2005) and streamflows (Hidalgo et al. 2009, Mantua et al. 2010, chapter 3922 3). There will likely be future declines in snowpack persistence and April 1 snow water 3923 equivalent throughout the Pacific Northwest, with the largest declines in mid-elevation, 3924 relatively wet locations (Luce et al. 2014). Extreme hydrologic events (e.g., those currently rated 3925 as having 100-year recurrence intervals) may become more frequent with future increases in 3926 temperature and potential increases in the amount of precipitation in winter months (Hamlet et 3927 al. 2013). For Southwest Washington, analyses indicate that the number of days with high 3928 streamflows in winter could increase 20-45 percent, and the magnitude of peak flows could 3929 increase 10-23 percent (Safeeq et al. 2015, chapter 3). Increased peak flows would affect erosion 3930 and sedimentation, which could, in turn, affect channel form and the fluvial dynamics of streams 3931 and their riparian zones (Capon et al. 2013). 3932 Summer streamflows may decrease with warming climate because of earlier snowmelt 3933 and earlier dates for peak streamflows (Leppi et al. 2011, Luce and Holden 2009, Safeeq et al. 3934 2013, Stewart et al. 2005). In Southwest Washington, mean summer streamflows may decrease 3935 40-57 percent (chapter 3). In addition, increased temperatures may lead to increased frequency 3936 and severity of droughts and increased area burned by wildfire (Littell et al. 2010, 2013). 3937 Increasing temperatures and evapotranspiration and decreasing summer streamflows may also 3938 lead to drying in some riparian areas, particularly on the eastside of the Cascade Range (Dwire 3939 and Mellmann-Brown 2017). Drying in riparian areas could decrease the extent of the riparian 3940 zone in some locations and result in shifts in riparian plant community composition. 3941 Cold-water corydalis and other species that rely specifically on cold, flowing water are 3942 particularly vulnerable to warming and drying in riparian areas. Shifts in riparian vegetation will 3943 depend on elevation, location within a watershed, and land use. However, shifts to more drought- 3944 tolerant species can be expected, and shifts to more disturbance-tolerant species, such as red 3945 alder, may occur with increased flooding, wildfire, and insect outbreaks. Nonnative species may 3946 also become more competitive in riparian areas with increased opportunities for invasion after 3947 disturbance (Catford et al. 2013). Changes in riparian plant species composition and reduced 3948 riparian extent could result in direct losses to the quantity and quality of ecological contributions 3949 of riparian vegetation, such as wildlife habitat, shade over streams, and buffer capacity for 3950 maintenance of water quality (Capon et al. 2013, Dwire and Mellmann-Brown 2017). Loss of 3951 riparian trees (e.g., willows [Salix spp.]) would also reduce an important habitat element for 3952 American beavers (Castor canadensis L.), which would, in turn, reduce retention of cool water 3953 behind beaver dams. 3954 3955 3956 Wetlands and Groundwater-Dependent Ecosystems 3957 3958 Wetlands are ecosystems “that are inundated or saturated by surface or ground water at a 3959 frequency and duration sufficient to support, and that under normal circumstances do support, a 3960 prevalence of vegetation typically adapted for life in saturated soil conditions” (FICWD 1989). 3961 Three broad categories of wetlands occur in Southwest Washington: lacustrine, riverine, and 3962 palustrineDRAFT (Cowardin et al. 1979). Lacustrine wetlands are associated with lakes or other bodies 92

Hudec et al. Chapter 5 93

3963 of water. Riverine wetlands are associated with flowing water. Palustrine wetlands are freshwater 3964 wetlands that are non-tidal and are not associated with flowing water. 3965 Riverine wetlands dominated by western redcedar (Thuja plicata Donn ex D. Don) are of 3966 special interest in GPNF. These low-elevation, valley-bottom wetlands are typically found on the 3967 east side of GPNF in the Little White Salmon River, Cave Creek, and White Salmon River 3968 drainages (Topik 1989). Mature forests in these western redcedar wetlands have high species 3969 diversity and provide a variety of wildlife habitat structures and food sources. Western redcedar 3970 is also a common component of small depressional wetlands, or forested ephemeral “pothole 3971 wetlands”, scattered throughout dry forests on the east side of GPNF. Pothole wetlands provide 3972 water sources for wildlife in these dry areas. 3973 Fens are palustrine wetlands supported primarily by groundwater with a minimum depth 3974 of 30-40 cm of accumulated peat (Chadde et al. 1998, USDA FS 2012). Fens are often 3975 dominated by herbaceous vegetation, but they may also be dominated by shrubs or trees. A 2010 3976 survey documented 14 fens in GPNF, with the majority on the Mt. Adams Ranger District at 3977 900-1,400 m elevation (Dewey 2011). Most fens in GPNF are parts of wetland complexes that 3978 include fen and wet meadow elements and sometimes dry meadow elements (Dewey 2011). 3979 These complexes providing habitat for a diversity of plant and animal species, including 3980 botanical species such as bladderwort (Utricularia spp.). Grand Meadows, Swampy Meadows, 3981 McClellan Meadows, and South Prairie Bog (actually a fen) are examples of wetland complexes 3982 with diverse plant communities and sensitive botanical species. 3983 Wetlands in Southwest Washington are expected to be highly vulnerable to climate 3984 change (Lee et al. 2015) as a function of altered snowpack (Hamlet et al. 2005), precipitation 3985 regimes (Luce et al. 2012, 2013; Safeeq et al. 2013), and groundwater recharge and discharge 3986 (Waibel et al. 2013). Warming in all seasons and reduced summer precipitation (Mote et al. 3987 2013) would result in increased evapotranspiration, increased soil moisture stress in summer 3988 (Littell et al. 2013), earlier drawdown, a more rapid recession rate in summer, and reduced 3989 minimum water levels in wetlands (Lee et al. 2015). 3990 Many wetlands are groundwater-dependent, and snowpack is the main source of 3991 groundwater recharge in montane areas (Winograd et al. 1998). Reduced snowpack with climate 3992 change would decrease the length of time aquifer recharge can occur, potentially leading to faster 3993 runoff, less groundwater recharge, and less groundwater to support springs and groundwater- 3994 dependent wetlands (Dwire and Mellmann-Brown 2017). Some groundwater-dependent 3995 wetlands may decrease in size or completely dry out in summer. However, effects will vary 3996 depending on hydrogeologic setting (Drexler et al. 2013). Some groundwater resources may be 3997 less sensitive to climate change than surface water, depending on local and regional geology, and 3998 surrounding land and water use (Tague and Grant 2009). Slowly infiltrating precipitation that 3999 includes both rain and snow could recharge groundwater aquifers as effectively as rapid, 4000 seasonal snowmelt runoff (Dwire and Mellmann-Brown 2017). 4001 Peat accumulation is a key ecological process in fens, and this process will likely be 4002 affected by climate change through higher temperatures and altered hydrology, such as lowered 4003 groundwater levels. These changes could lead to soil cracking, peat subsidence, and secondary 4004 alterations in water flow and storage patterns in peatlands with lowered groundwater (Kværner 4005 and Snilsberg 2011). Plant species that inhabit fens are highly sensitive to changes in water table 4006 elevation (Magee and Kentula 2005, Shipley et al. 1991), and compositional shifts would likely 4007 occur in fensDRAFT with lower water tables. Drought-tolerant species may become more competitive, 93

Hudec et al. Chapter 5 94

4008 rare and sensitive species could be extirpated, and cover of nonnatives may increase (Dwire and 4009 Mellmann-Brown 2017). 4010 Ephemeral or intermediate wetlands at higher elevations are expected to be highly 4011 sensitive to a warmer climate; some ephemeral montane wetlands may disappear, and some 4012 intermediate montane wetlands may become ephemeral (Lee et al. 2015). Some perennial 4013 montane wetlands in Washington (Olympic Peninsula and Cascade Range) may transition to 4014 intermediate wetlands or even to ephemeral wetlands as wetland water levels decline. Wetlands 4015 at lower elevations will be vulnerable to increasing water demands, pressure for increased 4016 diversion or water development, and other land-use activities that require water (Dwire and 4017 Mellmann-Brown 2017). 4018 Climate change may affect insects and amphibians that inhabit wetlands. Odonate 4019 (dragonfly, damselfly) nymphs can be top predators in fishless wetlands and are an essential food 4020 resource in wetland with fish and amphibians. Adult odonates are often eaten by upland 4021 predators such as birds, bats, and lizards. Warmer temperatures may affect development, 4022 phenology, behavior, and other characteristics of odonates (WDFW 2015). 4023 Oregon spotted frog (Rana pretiosa Baird and Girard) (Washington State endangered and 4024 federally threatened species) exists in six river drainages in Washington, including Trout Lake 4025 Creek in the southeast portion of GPNF and Outlet Creek in Conboy Lake National Wildlife 4026 Refuge, southeast of GPNF (Hallock 2013). Oregon spotted frog populations are often 4027 geographically isolated, and population numbers can fluctuate widely. Life history of the frog is 4028 completely aquatic, associated with relatively large permanent wetlands (typically greater than 4 4029 ha) connected to shallow, warm-water breeding habitats. A warmer climate may negatively 4030 impact Oregon spotted frog as seasonal drying of aquatic habitats and subsequent changes in 4031 vegetation (expansion of trees, shrubs, and nonnative species) become more common (WDFW 4032 2015). However, increased disturbance could improve breeding conditions for Oregon spotted 4033 frogs, because the species relies on seasonally flooded areas of shallow water with short 4034 vegetation and full sun exposure, which is typical of early-seral wetland plant growth. 4035 4036 4037 Conclusions 4038 4039 The special habitats discussed here are “special,” at least in part, because although they represent 4040 a small portion of the Southwest Washington landscape, they contain a high component of 4041 biological diversity, including many species of concern. Therefore, potential climate change 4042 effects on these special habitats may alter the distribution and abundance of species, as well as 4043 ecosystem structure and function. In some cases, particularly in systems dependent on specific 4044 hydrologic regimes, special habitats may already be experiencing impacts of climate change, and 4045 these impacts may be difficult to reverse or even mitigate in the future. 4046 Ecological disturbance and extreme climate-related events are expected to be the main 4047 transformational agents in forested systems associated with climate change. Increasing frequency 4048 and magnitude of droughts will create stress for trees and other species in summer when water 4049 demand is high and water supply is typically low. This will reduce growth and vigor in forests, 4050 making some species more susceptible to secondary stressors such as insects. If insect outbreaks 4051 become more chronic, they may kill old trees with high habitat value in drier portions of the 4052 landscape. Increased drought will also facilitate more frequent wildfire, in some cases burning 4053 areas that DRAFTrarely experience fire (e.g., westside late-successional forest). Tree mortality related to 94

Hudec et al. Chapter 5 95

4054 insects and fire may lead to a more fragmented forest landscape with a greater proportion of 4055 early-seral habitat than currently exists in some locations. 4056 The biggest near-term effects of climate change will probably occur in riparian areas, 4057 wetlands, and groundwater-dependent ecosystems. Reduced snowpack associated with warmer 4058 temperatures is already occurring in western Washington, with associated changes in hydrologic 4059 systems—primarily higher winter streamflows and lower summer streamflows. Less water 4060 during the normal summer dry period is especially stressful in water-dependent systems, and 4061 over time, can be expected to alter habitat for both plants and animals. Effects of altered 4062 hydrologic regimes will likely occur first in smaller habitats such as wet meadows and ephemeral 4063 ponds and streams. Few adaptation options are available to mitigate climate change impacts, 4064 although maintaining viable beaver populations can enhance retention of (cool) water where 4065 trees are available for building dams and lodges. Monitoring of riparian, wetland, and 4066 groundwater-dependent habitats in GPNF will be critical for tracking climate-related changes. 4067 4068 4069 Literature Cited 4070 4071 Agee, J.K. 1990. The historical role of fire in Pacific Northwest forests. In: Walstad, J.D.; 4072 Radosevich, S.R.; Sandberg, D.V., eds. Natural and prescribed fire in Pacific Northwest 4073 forests. Corvallis, OR: Oregon State University Press: 25–38. 4074 Agee, J.K. 1993. Fire ecology of Pacific Northwest forests. Covelo, CA: Island Press. 4075 Applegarth, J. 1995. Invertebrates of special status or special concern in the Eugene District. 4076 Unpublished report. Eugene, OR: U.S. Department of Interior, Bureau of Land Management. 4077 Arno, S.F. 1986. Whitebark pine cone crops: a diminishing source of wildlife food. Western 4078 Journal of Applied Forestry. 9: 92–94. 4079 Atkins, J. 2012. Conservation status of the Cascade red fox. Sci. Brief. Ashford, WA: U.S. 4080 Department of Interior, National Park Service, Mount Rainier National Park. 4081 Aubry, C.; Devine, W.; Shoal, R. [et al.]. 2011. Climate change and forest biodiversity: a 4082 vulnerability assessment and action plan for national forests in western Washington. 4083 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Region. 4084 Aubry, K.B.; Houston, D.B. 1992. Distribution and status of the fisher (Martes pennanti) in 4085 Washington. Northwest Naturalist. 73: 69–79. 4086 Aubry, K.B.; Lewis, J.C. 2003. Extirpation and reintroduction of fishers (Martes pennanti) in 4087 Oregon: implications for their conservation in the Pacific states. Biological Conservation. 4088 114: 79–90. 4089 Beniston, M. 2003. Climatic change in mountain regions: a review of possible impacts. Climatic 4090 Change. 59: 5–31. 4091 Betts, M.G.; Hagar, J.C.; Rivers, J.W. [et al.]. 2010. Thresholds in forest bird occurrence as a 4092 function of the amount of early-seral broadleaf forest at landscape scales. Ecological 4093 Applications. 20: 2116–2130. 4094 Brockman, C. 1958. Golden chinkapin (Castanopsis chrysophylla [Hook.] DC.). University of 4095 Washington Arboretum Bulletin. 21: 46–47, 70. 4096 Buchanan, J.B. 2015. Draft periodic status review for the northern spotted owl in Washington. 4097 Olympia, WA: Washington Department of Fish and Wildlife. 4098 Buchanan, J.B.; Lundquist, R.W.; Aubry, K.B. 2013. Winter populations of Douglas squirrels 4099 in differentDRAFT-aged Douglas-fir forests. Journal of Wildlife Management. 54: 577–581. 95

Hudec et al. Chapter 5 96

4100 Burke, T. 1999. Management recommendations for terrestrial mollusk species Cryptomastix 4101 devia, the Puget Oregonian. V. 2.0. Portland, OR: U.S. Department of Agriculture, Forest 4102 Service and U.S. Department of Interior, Bureau of Land Management. 4103 http://www.blm.gov/or/plans/surveyandmanage/MR/TM4Species/2000-015_1.pdf. (10 4104 February 2017). 4105 Burke, T.E. 1994. Survey of the Taneum watershed for species of the phylum Mollusca. Unpub. 4106 rep. Cle Elum, WA: U.S. Department of Agriculture, Forest Service, Okanogan-Wenatchee 4107 National Forest. 4108 Burrows, C.J. 1990. Processes of vegetation change. London, United Kingdom: Unwin Hyman 4109 Publishing. 4110 Capon, S.J.; Chambers, L.E.; Mac Nally, R. [et al.]. 2013. Riparian ecosystems in the 21st 4111 century: hotspots for climate change adaptation? Ecosystems. 16: 359–381. 4112 Carr, D.; Bowman, J.; Wilson, P.J. 2007. Denisty-dependent dispersal suggests a genetic 4113 measure of habitat suitability. Oikos. 116: 629–635. 4114 Carroll, C. 2010. Role of climatic niche models in focal-species-based conservation planning: 4115 assessing potential effects of climate change on northern spotted owl in the Pacific 4116 Northwest, USA. Biological Conservation. 143: 1432–1437. 4117 Carroll, C.; Zielinski, W.J.; Noss, R.F. 1999. Using presence-absence data to build and test 4118 spatial habitat models for the fisher in the Klamath region. Conservation Biology. 13: 1344– 4119 1359. 4120 Case, M. 2009. Ursus americanus-Olympics. Pacific Northwest climate change vulnerability 4121 assessment. Climate sensitivity database. http://www.climatevulnerability.org. (10 February 4122 2017). 4123 Catford, J.A.; Naiman, R.J.; Chambers, L.E. 2013. Predicting novel riparian ecosystems in a 4124 changing climate. Ecosystems. 16: 382–400. 4125 Chadde, S.W.; Shelly, J.S.; Bursik, R.J. [et al.]. 1998. Peatlands on national forests of the 4126 Northern Rocky Mountains: ecology and conservation. Gen. Tech. Rep. RMRS-GTR-11. 4127 Ogden, UT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 4128 Station. 4129 Chappell, C.B. 2006. Plant associations of balds and bluffs of western Washington. Natur. 4130 Heritage Rep. 2006-02. Olympia, WA: Washington Department of Natural Resources. 4131 Chappell, C.B.; Crawford, R.C; Barrett, C. [et al.]. 2001. Wildlife habitats: descriptions 4132 status, trends, and system dynamics. In: Johnson, D.H.; O’Neil, T.A., eds. Wildlife habitat 4133 relationships in Oregon and Washington. Corvallis, OR: Oregon State University Press. 4134 Chen, I.C.; Hill, J.K.; Ohlemueller, R. [et al.]. 2011. Rapid range shifts of species associated 4135 with high levels of climate warming. Science. 333: 1024–1026. 4136 Copeland, J.P.; McKelvey, K.S.; Aubry, K.B. [et al.]. 2010. The bioclimatic envelope of the 4137 wolverine (Gulo gulo): do climatic constraints limit its geographic distribution? Canadian 4138 Journal of Zoology. 88: 233–246. 4139 Cowardin, L.M.; Carter, V.; Golet, F.C.; LaRoe, E.T. 1979. Classification of wetlands and 4140 deepwater habitats of the United States. FWS/OBS-79/31. Washington, DC: U.S. 4141 Department of the Interior, U.S. Fish and Wildlife Service. 4142 Davis, F.W.; Seo, C.; Zielinski, W.J. 2007. Regional variation in home-range-scale habitat 4143 models for fisher (Martes pennanti) in California. Ecological Applications. 17: 2195–2213. 4144 Davis, R.J.; Ohmann, J.L.; Kennedy, R.E. [et al.]. 2015. Northwest Forest Plan–the first 20 4145 years (1994DRAFT-2013): status and trends of late-successional and old-growth forests. PNW-GTR- 96

Hudec et al. Chapter 5 97

4146 911. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 4147 Research Station. 4148 Dewey, R. 2011. Detection and bryological inventory of rich fens on Gifford Pinchot National 4149 Forest. https://www.fs.fed.us/r6/sfpnw/issssp/documents/inventories/inv-rpt-br-gip-fen- 4150 inventory-2011-01.pdf. (8 February 2018). 4151 Donato, D.C.; Campbell, J.L.; Franklin, J.F. 2012. Multiple successional pathways and 4152 precocity in forest development: can some forests be born complex? Journal of Vegetation 4153 Science. 23: 576–584. 4154 Drexler, J.A., Knifong, D.; Tuil, J. [et al.]. 2013. Fens as whole-ecosystem gauges of 4155 groundwater recharge under climate change. Journal of Hydrology. 481: 22–34. 4156 Dwire, K.A.; Mellmann-Brown, S. 2017. Climate change and special habitats in the Blue 4157 Mountains: riparian areas, wetlands, and groundwater-dependent ecosystems. In: Halofsky, 4158 J.E.; Peterson, D.L., eds. Climate change vulnerability and adaptation in the Blue Mountains. 4159 PNW-GTR-939. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 4160 Northwest Research Station. 4161 Farnes, P.E. 1990. Snotel and snow course data: describing the hydrology of whitebark 4162 pine ecosystems. In: Schmidt, W.C.; McDonald, K.J., comps. Proceedings— 4163 symposium on whitebark pine ecosystems: ecology and management of a high- 4164 mountain resource; 1989 March 29-31; Bozeman, MT. Gen. Tech. Rep. INT-270. 4165 Ogden, UT: U.S. Department of Agriculture, Intermountain Research Station: 302– 4166 304. 4167 Federal Interagency Committee for Wetland Delineation [FICWD] 1989. Federal manual for 4168 identifying and delineating wetlands. Coop. Tech. Pub. Washington, DC: Department of 4169 Defense, U.S. Army Corps of Engineers; U.S. Environmental Protection Agency; 4170 Department of the Interior, U.S. Fish and Wildlife Service; U.S. Department of Agriculture, 4171 Soil Conservation Service. 4172 Forsman, E.D.; Anthony, R.G.; Dugger, K.M. [et al.]. 2011. Population demography of 4173 northern spotted owls: 1985-2008. Studies in Avian Biology 40. Chicago, IL: American 4174 Ornithological Society. 4175 Frey, B.R.; Lieffers, V.J.; Hogg, E.H., Landhäusser, S.M. 2004. Predicting landscape patterns 4176 of aspen dieback: mechanisms and knowledge gaps. Canadian Journal of Forest Research. 4177 34: 1379–1390. 4178 Friggens, M.M.; Williams, M.; Bagne, K. [et al.]. 2018. Effects of climate change on 4179 terrestrial animals. In: Halofsky, J.E.; Peterson, D.L.; Ho, J.J.; Little, N., eds. Climate change 4180 vulnerability and adaptation in the Intermountain region. Gen. Tech. Rep. RMRS-GTR- 4181 XXX. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain 4182 Research Station. 4183 Geils, B.W.; Hummer, K.E.; Hunt, R.S. 2010. White pines, Ribes, and blister rust: a review 4184 and synthesis. Forest Pathology. 40:147–185. 4185 Gregory, S.V.; Swanson, F.J.; McKee, W.A.; Cummins, K.W. 1991. An ecosystem 4186 perspective of riparian zones. Bioscience. 41: 540–551. 4187 Gutiérrez, R.J.; Franklin, A.B.; Lahaye, W.S. 1995. Spotted owl (Strix occidentalis). In: 4188 Poole, A., ed. The birds of North America online Ithaca, NY: Cornell Laboratory of 4189 Ornithology. 4190 Hallock, L.A. 2013. Draft Washington state Oregon spotted frog recovery plan. Olympia, WA: 4191 WashingtonDRAFT Department of Fish and Wildlife. 97

Hudec et al. Chapter 5 98

4192 Halofsky, J.E.; Piper, S.; Aluzas, K. [et al.]. 2011. Climate change, wildlife management, and 4193 habitat management at Olympic National Forest and Olympic National Park. In: Halofsky, 4194 J.E.; Peterson, D.L.; O’Halloran, K.A.; Hawkins Hoffman, C., eds. Adapting to climate 4195 change at Olympic National Forest and Olympic National Park. PNW-GTR-844. Portland, 4196 OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station: 4197 91–118. 4198 Hamer, T.E. 1988. Home range size of the northern barred owl and northern spotted owl in 4199 western Washington. Master's thesis. Bellingham, WA: Western Washington University. 4200 Hamer, T.E.; Nelson S.K., 1995. Characteristics of marbled murrelet nest trees and nesting 4201 stands. In: Ralph, C.; Hunt, G.L.; Raphael, M.G. Piatt, J.F., eds. Ecology and conservation of 4202 the marbled murrelet. PSW-GTR-152. U.S. Department of Agriculture, Forest Service, 4203 Pacific Southwest Research Station: 69–82. 4204 Hamlet, A.F., Elsner, M.M.; Mauger, G.S. [et al.]. 2013. An overview of the Columbia Basin 4205 climate change scenarios project: approach, methods, and summary of key results. 4206 Atmosphere-Ocean. 51: 392–415. 4207 Hamlet, A.F.; Mote, P.W.; Clark, M.P.; Lettenmaier, D.P. 2005. Effects of temperature and 4208 precipitation variability on snowpack trends in the western U.S. Journal of Climate. 18: 4209 4545–4561. 4210 Hannah, L., Midgely, G.; Hughs, G.; Bomhard, B. 2005. The view from the Cape: extinction 4211 risk, protected areas, and climate change. BioScience. 55: 231–242. 4212 Harvey, D.S.; Weatherhead, P.J. 2010. Habitat selection as the mechanism for 4213 thermoregulation in a northern population of massasauga rattlesnakes (Sistrurus catenatus). 4214 Ecoscience. 17: 411–419. 4215 Hessburg, P.F.; Agee, J.K.; Franklin, J.F. 2005. Dry forests and wildland fires of the inland 4216 Northwest USA: contrasting the landscape ecology of the pre-settlement and modern eras. 4217 Forest Ecology and Management. 211: 117–139. 4218 Hidalgo, H.G.; Das, T.; Dettinger, M.D. [et al.]. 2009. Detection and attribution of streamflow 4219 timing changes to climate change in the western United States. Journal of Climate. 22: 3838– 4220 3855. 4221 Holms, T.H. 1990. Botanical trends in northern California oak woodlands. Rangelands. 12: 3–7. 4222 Howard, J.L. 1996. Populus tremuloides. In: Fire Effects Information System. U.S. Department 4223 of Agriculture, Forest Service. Rocky Mountain Research Station. 4224 http://www.fs.fed.us/database/feis. (10 February 2017). 4225 Ingram, R. 1973. Wolverine, fisher, and marten in central Oregon. Central Reg. Admin. Rep. 4226 Salem, OR: Oregon State Game Commission. 4227 Inkley, D.B.; Anderson, M.G.; Blaustein, A.R. [et al.]. 2004. Global climate change and 4228 wildlife in North America. Tech. Review 04-2. Bethesda, MD: The Wildlife Society. 4229 Inman, R.M.; Magoun, A.J.; Persson, J.; Mattisson, J. 2012. The wolverine’s niche: linking 4230 reproductive chronology, caching, competition, and climate. Journal of Mammalogy. 93: 4231 634–644. 4232 Jacobson, J.E.; Lewis, J.C.; Snyder, M.C. 2003. Assessment of fisher habitat in Washington 4233 State. Tier 1 refinement and tier 2 final rep. Olympia, WA: Washington Department of Fish 4234 and Wildlife. 4235 Jakubowski, J. 2015. Mardon skipper (Polites mardon) site management plans. Vancouver, 4236 WA: U.S.DRAFT Department of Agriculture, Forest Service, Gifford Pinchot National Forest. 98

Hudec et al. Chapter 5 99

4237 http://www.fs.fed.us/r6/sfpnw/issssp/documents3/smp-iile-polites-mardon-gip-cvrd-2015-10- 4238 508.pdf. (10 February 2017). 4239 Jeffress, M.R.; Rodhouse, T.J.; Ray, C. [et al.]. 2013. The idiosyncrasies of place: Geographic 4240 variation in the climate-distribution relationships of the American pika. Ecological 4241 Applications. 23: 864–878. 4242 Jencso, K.G.; McGlynn, B.L. 2011. Hierarchical controls on runoff generation: 4243 Topographically driven hydrologic connectivity, geology, and vegetation. Water Resources 4244 Research. 47: W11527. 4245 Johnson, T.E.; Butcher, J.; Parker, A.; Weaver, C.P. 2012. Investigating the implications of 4246 climate change for U.S. stream water quality: the EPA Global Change Research Program’s 4247 “20 Watersheds” project. Journal of Water Resources Planning and Management. 138: 453– 4248 464. 4249 Johnston, K.M.; Freund, K.A.; Schmitz, O.J. 2012. Projected range shifting by montane 4250 mammals under climate change: implications for Cascadia's National Parks. Ecosphere. 3: 1– 4251 51. 4252 Jones, J.R.; DeByle, N.Y. 1985. Aspen, ecology and management in the western United States. 4253 RM-GTR-119. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 4254 Mountain Forest and Range Experiment Station. 4255 Keane, R.E.; Tomback, D.F.; Aubry, C.A. [et al.]. 2012. A range-wide restoration strategy for 4256 whitebark pine (Pinus albicaulis). Gen. Tech. Rep. RMRS-GTR-279. Fort Collins, CO: U.S. 4257 Department of Agriculture, Forest Service, Rocky Mountain Research Station. 4258 Keeler-Wolf, T. 1988. The role of Chrysolepis chrysophylla (Fagaceae) in the Pseudotsuga 4259 hardwood forest of the Klamath Mountains of California. Madroño. 35: 285–308. 4260 Kertis, J. 1986. Vegetation dynamics and disturbance history of Oak Patch Preserve, Mason 4261 County, Washington. Unpub. rep. Olympia, WA: Washington Department of Natural 4262 Resources. 4263 Körner, C. 1998. Worldwide positions of alpine treelines and their causes. In: Beniston, M.; 4264 Innes, J.L.; eds. The impacts of climate variability on forests. Heidelberg, Germany: 4265 Springer-Verlag: 221–229. 4266 Krohn, W.B.; Elowe, K.D.; Boone, R.B. 1995. Relations among fishers, snow, and martens: 4267 Development and evaluation of two hypotheses. Forestry Chronicle. 71: 97–105. 4268 Kruckeberg, A.R. 1980. Golden chinquapin (Chrysolepis chrysophylla) in Washington state: a 4269 species at the northern limit of its range. Northwest Science. 54: 9–16. 4270 Kværner, J.; Snilsberg, P. 2011. Groundwater hydrology of boreal peatlands above a bedrock 4271 tunnel—drainage impacts and surface water groundwater interactions. Journal of Hydrology. 4272 403: 278–291. 4273 Larsen, E.M.; Morgan, J.T. 1998. Management recommendations for Washington’s priority 4274 habitats: Oregon white oak woodlands. Olympia, WA: Washington Department of Fish and 4275 Wildlife. 4276 Lawler, J.J.; Raymond, C.L.; Ryan, M.E. [et al.]. 2015. Climate change, wildlife, and wildlife 4277 habitat in the North Cascade Range. In: Raymond, C.L.; Peterson, D.L.; Rochefort, R.M, eds. 4278 Climate change vulnerability and adaptation in the North Cascades region, Washington. Gen. 4279 Tech. Rep. PNW-GTR-892. Portland, OR: U.S. Department of Agriculture, Forest Service, 4280 Pacific Northwest Research Station: 177–234. 4281 Lawler, J.J.; Safford, H.D.; Girvetz, E.H. 2012. Martens and fishers in a changing climate. In: 4282 Aubry,DRAFT K.B.; Zielinski, W.J.; Raphael, M.G. [et al.], eds. Biology and conservation of 99

Hudec et al. Chapter 5 100

4283 martens, sables, and fishers: a new synthesis; Ithaca, NY: Comstock Publications Associates: 4284 371–397. 4285 Lee, S-Y; Ryan, M.E.; Hamlet, A.F., [et al.]. 2015. Projecting the hydrologic impacts of 4286 climate change on montane wetlands. PLoS ONE. 10: e0136385. 4287 Leibowitz, S.G.; Comeleo, R.L.; Wigington, P.J. [et al.]. 2014. Hydrologic landscape 4288 classification evaluates streamflow vulnerability to climate change in Oregon, USA. 4289 Hydrology and Earth System Sciences. 18: 3367–3392. 4290 Leppi, J.C.; DeLuca, T.H.; Harrar, S.W.; Running, S.W. 2011. Impacts of climate change on 4291 August stream discharge in the Central-Rocky Mountains. Climatic Change. 112: 997–1014. 4292 Leventhal, R. 2007. The hoary marmot of the North Cascades in western Washington. Master’s 4293 thesis. Bellingham, WA: Western Washington University. 4294 Lewis, J.C.; Hayes, G.E. 2004. Feasibility assessment for reintroducing fishers to Washington. 4295 Olympia, WA: Department of Fish and Wildlife. 4296 Lewis, J.C.; Powell, R.A.; Zielinski, W.J. 2012. Carnivore translocations and conservation: 4297 insights from population models and field data for fishers (Martes pennanti). PLoS One 7: 4298 e32726. 4299 Lewis, J.C.; Stinson, D.W. 1998. Washington state status report for the fisher. Olympia, WA: 4300 Department of Fish and Wildlife. 4301 Littell, J.S.; Hicke, J.A.; Shafer, S.L. [et al.]. 2013. Forest ecosystems—vegetation, 4302 disturbance, and economics. In: Dalton, M.M.; Mote, P.W.; Snover, A.K., eds. 2013. Climate 4303 change in the Northwest: implications for our landscapes, waters, and communities. 4304 Washington, DC: Island Press: 110–148. 4305 Littell, J.S.; Oneil, E.E.; McKenzie, D. [et al.]. 2010. Forest ecosystems, disturbance, and 4306 climatic change in Washington State, USA. Climatic Change. 102: 129–159. 4307 Loehman, R.A.; Corrow, A.; Keane, R.E. 2011. Modeling climate changes and wildfire 4308 interactions: effects on whitebark pine (Pinus albicaulis) and implications for restoration, 4309 Glacier National Park, MT. In: Keane, R.E.; Tomback, D.F.; Murray, M.P.; Smith, C.M., 4310 eds. The future of high-elevation, five-needle white pines in Western North America: 4311 Proceedings of the High Five Symposium. 28-30 June 2010, Missoula, MT. Proc. RMRS-P- 4312 63. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain 4313 Research Station: 176–188. 4314 Lofroth, E.C.; Raley, C.M.; Finley, L.L.; Naney, R.H. 2010. Conservation of fishers (Martes 4315 pennanti) in south-central British Columbia, western Washington, western Oregon, and 4316 California. Denver, CO: Bureau of Land Management. 4317 Luce, C.; Morgan, P.; Dwire, K. [et al.]. 2012. Climate change, forests, fire, water, and fish: 4318 building resilient landscapes, streams, and managers. Gen. Tech. Rep. RMRS-GTR-290. Fort 4319 Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 4320 Station. 4321 Luce, C.H.; Abatzoglou, J.T.; Holden, Z.A. 2013. The missing mountain water: slower 4322 westerlies decrease orographic enhancement in the Pacific Northwest USA. Science. 342: 4323 1360–1364. 4324 Luce, C.H.; Holden, Z.A. 2009. Declining annual streamflow distributions in the Pacific 4325 Northwest United States, 1948–2006. Geophysical Research Letters. 36: L16401. 4326 Luce, C.H.; Lopez-Burgos, V.; Holden, Z. 2014. Sensitivity of snowpack storage to 4327 precipitation and temperature using spatial and temporal analog models. Water Resources 4328 Research.DRAFT 50: 9447–9462. 100

Hudec et al. Chapter 5 101

4329 Magee, T.K; Kentula, M.E. 2005. Response of wetland plant species to hydrologic conditions. 4330 Wetlands Ecology and Management. 13: 163–181. 4331 Malanson, G.P.; Butler, D.R.; Fagre, D.B. [et al.]. 2007. Alpine treeline of western North 4332 America: linking organism-to-landscape dynamics. Physical Geography. 28: 378–396. 4333 Mantua, N.; Tohver, I.; Hamlet, A. 2010. Climate change impacts on streamflow extremes and 4334 summertime stream temperature and their possible consequences for freshwater salmon 4335 habitat in Washington State. Climatic Change. 102: 187–223. 4336 Marchetti, S.B.; Worrall, J.J.; Eager, T. 2011. Secondary insects and diseases contribute to 4337 sudden aspen decline in southwestern Colorado, USA. Canadian Journal of Forest Research. 4338 41: 2315–2325. 4339 Marks, R. 2006. Amphibians and reptiles. Fish and Wildlife Habitat Management Leaflet, No. 4340 35, February 2006. Natural Resources Conservation Service and Wildlife Habitat Council. 4341 https://www.whmi.nrcs.usda.gov/technical/leaflet.htm. (7 December 2017). 4342 McDonald, G.I.; Hoff, R.J. 2001. Blister rust: an introduced plaque. In: Tomback, D.F.; Arno, 4343 S.F.; Keane, R.E., eds. Whitebark pine communities: ecology and restoration. Washington, 4344 DC: Island Press: 193–220. 4345 McKee, A. 1990. Castanopsis chrysophylla (Dougl.) A. DC. Giant chinkapin. In: Burns, R.M.; 4346 Honkala, B.H., tech. coords. Silvics of North America. Vol. 2. Hardwoods. Agric. Handb. 4347 654. Washington, DC: U.S. Department of Agriculture, Forest Service: 234–239. 4348 McKelvey, K.S.; Buotte, P.C. 2018. Climate change and wildlife in the Northern Rocky 4349 Mountains. In: Halofsky, J.E.; Peterson, D.L.; Dante-Wood, K., eds. Climate change 4350 vulnerability and adaptation in the Northern Rocky Mountains. Gen. Tech. Rep. RMRS- 4351 GTR-374. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 4352 Mountain Research Station. 4353 McKinney, S.T.; Tomback, D.F. 2007. The influence of white pine blister rust on seed 4354 dispersal in whitebark pine. Canadian Journal of Forest Research 37: 1044–1057. 4355 Meyer, R. 2012. Chrysolepis chrysophylla. Fire Effects Information System. U.S. Department of 4356 Agriculture, Forest Service, Rocky Mountain Research Station. 4357 http://www.fs.fed.us/database/feis. (10 February 2017). 4358 Millar, C.I.; Westfall, R.D. 2010. Distribution and climatic relationships of the American pika 4359 (Ochotona princeps) in the Sierra Nevada and Western Great Basin, U.S.A.: periglacial 4360 landforms as refugia in warming climates. Arctic, Antarctic, and Alpine Research. 42: 76–88. 4361 Mote. P.; Abatzoglou, J.; Kunkel, K. 2013. Climate change and variability in the past and 4362 future. In: Dalton, M.M.; Mote, P.W.; Snover, A.K., eds. Climate change in the Northwest: 4363 implications for our landscapes, waters and communities. Washington, DC: Island Press: 25– 4364 40. 4365 Naiman, R.J., Décamps, H.; McCalin, M.E. 2005. Riparia: ecology, conservation, and 4366 management of streamside communities. Burlington, MA: Elsevier Academic Press. 4367 Naiman, R.J.; Décamps, H. 1997. The ecology of interfaces: riparian zones. Annual Review of 4368 Ecology and Systematics. 28: 621–658. 4369 Naiman, R.J.; Décamps, H. Pollock, M. 1993. The role of riparian corridors in maintaining 4370 regional biodiversity. Ecological Applications. 3: 209–212. 4371 Nelson, R.A.; Folk, G.E.; Pfeiffer, E.W. [et al.]. 1983. Behavior, biochemistry, and hibernation 4372 in black, grizzly, and polar bears. In: Bears: their biology and management, vol. 5, Fifth 4373 InternationalDRAFT Conference on Bear Research and Management: 284–290. 101

Hudec et al. Chapter 5 102

4374 Nelson, S.K. 1997. Marbled murrelet (Brachyramphus marmoratus). In Poole, A., ed. The birds 4375 of North America online. Ithaca, NY: Cornell Laboratory of Ornithology. 4376 Olson, G.S.; Anthony, R.G.; Forsman, E.D. [et al.]. 2005. Modeling of site occupancy 4377 dynamics for northern spotted owls, with emphasis on the effects of barred owls. Journal of 4378 Wildlife Management. 69: 918–932. 4379 Olson, L.E.; Sauder, J.D.; Albrecht, N.M. 2014. Modeling the effects of dispersal and patch 4380 size on predicted fisher (Pekania [Martes] pennanti) distribution in the U.S. Rocky 4381 Mountains. Biological Conservation. 169: 89–98. 4382 Parmesan, C. 2006. Ecological and evolutionary responses to recent climate change. Annual 4383 Review of Ecology Evolution and Systematics. 37: 637–669. 4384 Parmesan, C.; Yohe, G. 2003. A globally coherent fingerprint of climate change impacts across 4385 natural systems. Nature. 421: 37–42. 4386 Powers, J.S., Sollins, P., Harmon, M.E., Jones, J.A., 1999. Plant-pest interactions in time and 4387 space: a Douglas-fir bark beetle outbreak as a case study. Landscape Ecology. 14: 105–120. 4388 Powell, R.A. 1993. The fisher: life history, ecology, and behavior, St. Paul, MN: University of 4389 Minnesota Press. 4390 Purcell, K.L.; Mazzoni, A.K.; Mori, S.R.; Boroski, B.B. 2009. Resting structures and resting 4391 habitat of fishers in the southern Sierra Nevada, California. Forest Ecology and Management. 4392 258: 2696–2706. 4393 Pyle, R.M. 1989. Washington butterfly conservation status report and plan. Olympia, WA: 4394 Washington Department of Fish and Wildlife. 4395 Ray, C.; Saracco, J.F.; Holmgren, M.L. [et al.]. 2017. Recent stability of resident and 4396 migratory landbird populations in National Parks of the Pacific Northwest. Ecosphere. 8: 4397 e01902. 4398 Ray, E.J.; Bergey, E.A. 2015. After the burn: factors affecting land snail survival in post- 4399 prescribed-burn woodlands. Journal of Mollusk Studies. 81: 44–50. 4400 Raymondi, R.R.; Cuhaciyan, J.E.; Glick, P. [et al.]. 2013. Water resources. In: Dalton, M.M.; 4401 Mote, P.W.; Snover, A.K., eds. Climate change in the Northwest: implications for our 4402 landscapes, waters, and communities. Washington, DC: Island Press: 41–66. 4403 Rehfeldt, G.E.; Ferguson, D.E.; Crookston, N.L. 2009. Aspen, climate, and sudden decline in 4404 western USA. Forest Ecology and Management. 258: 2353–2364. 4405 Rochefort, R.M.; Peterson, D.L. 1996. Temporal and spatial distribution of trees in subalpine 4406 meadows of Mount Rainier National Park. Arctic and Alpine Research. 28: 52–59. 4407 Root, T.L.; Price, J.T.; Hall, K.R. [et al.]. 2003. Fingerprints of global warming on wild 4408 animals and plants. Nature. 421: 57–60. 4409 Ruchty, A. 2011. Conservation strategy for Sisyrinchium sarmentosum Suks. ex. Green. 4410 Portland, OR: U.S. Department of Agriculture, Forest Service and U.S. Department of 4411 Interior, Bureau of Land Management. 4412 Safeeq, M.; Grant, G.E.; Lewis, S.L.; Staab, B. 2015. Predicting landscape sensitivity to 4413 present and future floods in the Pacific Northwest, USA. Hydrological Processes. 29: 5337– 4414 5353. 4415 Safeeq, M.; Grant, G.E.; Lewis, S.L.; Tague, C.L. 2013. Coupling snowpack and groundwater 4416 dynamics to interpret historical streamflow trends in the western United States. Hydrological 4417 Processes. 27: 655–668. 4418 Schmitz, O.J.; Post, E.; Burns, C.E.; Johnston, K.M. 2003. Ecosystem responses to global 4419 climateDRAFT change: moving beyond color mapping. BioScience. 53: 1199–1205. 102

Hudec et al. Chapter 5 103

4420 Schneider, S.H.; Root, T.L.; Van Putten, M. 2002. Wildlife responses to climate change: 4421 North American case studies. Washington, DC: Island Press. 4422 Seager, R., Ting, M., Held, I. [et al.]. 2007. Model projections of an imminent transition to a 4423 more arid climate in southwestern North America. Science. 316: 1181–1184. 4424 Shepperd, W.D.; Rogers, P.C.; Burton, D.; Bartos, D.L. 2006. Ecology, biodiversity, 4425 management, and restoration of aspen in the Sierra Nevada. Gen. Tech. Rep. RMRS-GTR- 4426 178. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain 4427 Research Station. 4428 Shipley, B.; Keddy, P.A.; Lefkovich, L.P. 1991. Mechanisms producing plant zonation along a 4429 water depth gradient: a comparison with the exposure gradient. Canadian Journal of Botany. 4430 69: 1420–1424. 4431 Shoal, R. 2009. Occurrence and habitat status evaluation for golden chinquapin (Chrysolepis 4432 chrysophylla [Dougl. ex Hook.] Hjelmqvist) on the Olympic National Forest. Interag. Spec. 4433 Status Species Prog. Olympia, WA: U.S. Department of Agriculture, Forest Service, 4434 Olympic National Forest. 4435 Spencer, W.; Rustigian-Romsos, H.; Strittholt, J. [et al.]. 2011. Using occupancy and 4436 population models to assess habitat conservation opportunities for an isolated carnivore 4437 population. Biological Conservation. 144: 788–803. 4438 Stewart, I.T.; Cayan, D.R.; Dettinger, M.D. 2005. Changes toward earlier streamflow timing 4439 across western North America. Journal of Climatology. 18: 1136–1155. 4440 Sugihara, N.G.; Reed, L.J. 1987. Prescribed fire for restoration and maintenance of Bald Hills 4441 oak woodlands. In: Plumb, T.R.; Pillsbury, N.H., tech. coords. Proceedings of the symposium 4442 on multiple-use of California’s hardwood resources. Gen. Tech. Rep. PSW-GTR-100. 4443 Berkeley, CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research 4444 Station: 446–451. 4445 Swanson, M.E.; Franklin, J.F.; Beschta, R.L. [et al.]. 2011. The forgotten stage of forest 4446 succession: early-successional ecosystems on forest sites. Frontiers in Ecology and the 4447 Environment. 9: 117–125. 4448 Swanson, M.E.; Studevant, N.M.; Campbell, J.L.; Donato, D.C. 2014. Biological associates 4449 of early-seral pre-forest in the Pacific Northwest. Forest Ecology and Management. 324: 4450 160–171. 4451 Tague, C.; Grant, G.E. 2009. Groundwater dynamics mediate low-flow response to global 4452 warming in snow-dominated alpine regions. Water Resources Research. 45: W07421. 4453 Taylor, R.J.; Boss, T.R. 1975. Biosystematics of Quercus garryana in relation to its distribution 4454 in the state of Washington. Northwest Science. 49: 48–57. 4455 Tepley, A.J.; Swanson, F.J.; Spies, T.A. 2013. Fire-mediated pathways of stand development 4456 in Douglas-fir/western hemlock forests of the Pacific Northwest, USA. Ecology. 94: 1729– 4457 1743. 4458 Thilenius, J.F. 1968. The Quercus garryana forests of the Willamette Valley, Oregon. Ecology 4459 49: 1124–1133. 4460 Thomas, J.W.; Black, H.J.; Scherzinger, R.J.; Pedersen, R.J. 1979. Deer and elk. In: 4461 Thomas, J.W., tech. ed. Wildlife habitats in managed forests: the Blue Mountains of Oregon 4462 and Washington. Agric. Hndbk. 533. Washington, D.C.: U.S. Department of Agriculture: 4463 104–127.DRAFT 103

Hudec et al. Chapter 5 104

4464 Tomback, D.F. 2005. The impact of seed dispersal by the Clark’s nutcracker on whitebark pine: 4465 multi-scale perspective on a high mountain mutualism. In: Broll, G.; Kepline, B., eds. 4466 Mountain ecosystems: studies in treeline ecology. Berlin, Germany: Springer: 181–201. 4467 Tomback, D.F.; Arno, S.F.; Keane, R.E. 2001. The compelling case for management 4468 intervention. In: Tomback, D.F.; Arno, S.F.; Keane, R.E., eds. Whitebark pine 4469 communities: ecology and restoration. Washington, DC: Island Press: 3–25. 4470 Tomback, D.F.; Hoffman, L.A.; Sund, S.K. 1990. Coevolution of whitebark pine and 4471 nutcrackers: implications for forest regeneration. In: Schmidt, W.C.; McDonald, K.J., 4472 compilers. Proceedings—symposium on whitebark pine ecosystems: ecology and 4473 management of a high-mountain resource. Gen. Tech. Rep. GTR-INT-270. Ogden, UT: U.S. 4474 Department of Agriculture, Forest Service, Intermountain Research Station: 118–129. 4475 Tomback, D.F.; Kendall, K.C. 2001. Biodiversity losses: the downward spiral. In: 4476 Tomback, D.F.; Arno, S.F.; Keane, R.E., eds. Whitebark pine communities: ecology 4477 and restoration. Washington, DC: Island Press: 243–262. 4478 Topik, C. 1989. Plant association and management guide for the grand fir zone, Gifford Pinchot 4479 National Forest. R6-ECOL-TP_006-88. Portland, OR: U.S. Department of Agriculture, 4480 Forest Service, Pacific Northwest Region. 4481 U.S. Department of Agriculture, Forest Service [USDA FS]. 1990. Gifford Pinchot National 4482 Forest Land and Resource Management Plan. Vancouver, WA: U.S. Department of 4483 Agriculture, Forest Service, Gifford Pinchot National Forest. 4484 U.S. Department of Agriculture, Forest Service [USDA FS]. 1995. Inland native fish strategy. 4485 Environmental assessment: decision notice and finding of no significant impact. Portland, 4486 OR: U.S. Department of Agriculture, Forest Service, Intermountain, Northern, and Pacific 4487 Northwest Regions. 4488 U.S. Department of Agriculture, Forest Service [USDA FS]. 1997. Gifford Pinchot National 4489 Forest Forest-wide late-successional reserve assessment. Vancouver, WA: U.S. Department 4490 of Agriculture, Forest Service, Pacific Northwest Region. 4491 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012. National best management 4492 practices for water quality management on National Forest System lands. Vol. 1: national 4493 core BMP technical guide. Washington, DC: U.S. Department of Agriculture, Forest Service. 4494 U.S. Department of Agriculture, Forest Service; U.S. Department of Interior, Bureau of 4495 Land Management [USDA FS and USDI BLM]. 1994. Northwest Forest Plan record of 4496 decision for amendments to Forest Service and Bureau of Land Management planning 4497 documents within the range of the northern spotted owl. Portland, OR: U.S. Department of 4498 Agriculture, Forest Service; U.S. Department of Interior, Bureau of Land Management. 4499 U.S. Department of Interior, National Park Service [USDI NPS]. 2016. Meadows-Yosemite 4500 National Park. http://www.nps.gov/yose/learn/nature/meadows.htm. (10 February 2017). 4501 Waibel, M.S.; Gannett, M.W.; Chang, H.; Hulbe, C.L. 2013. Spatial variability of the 4502 response to climate change in regional groundwater systems—examples from simulations in 4503 the Deschutes Basin, Oregon. Journal of Hydrology. 486: 187–201. 4504 Waring, K.M.; Six, D.L. 2005. Distribution of bark beetle attacks after whitebark pine 4505 restoration treatments: a case study. Western Journal of Applied Forestry. 20: 110–116. 4506 Washington Department of Fish and Wildlife [WDFW]. 2013. Western gray squirrels. 4507 http://wdfw.wa.gov/conservation/gray_squirrel/graphics/maps/finalwesterngraysquirrel.jpg. 4508 (10 FebruaryDRAFT 2017). 104

Hudec et al. Chapter 5 105

4509 Washington Department of Fish and Wildlife [WDFW]. 2015. Washington’s State Wildlife 4510 Action Plan, 2015 update. Olympia, WA: Washington Department of Fish and Wildlife. 4511 White, K.S.; Pendleton, G.W.; Crowley, D. [et al.]. 2011. Mountain goat survival in coastal 4512 Alaska: effects of age, sex and climate. Journal of Wildlife Management. 75: 1731–1744. 4513 Wiens, J.D.; Anthony, R.G.; Forsman, E.D. 2014. Competitive interactions and resource 4514 partitioning between northern spotted owls and barred owls in western Oregon. Wildlife 4515 Monographs. 185: 1–50. 4516 Winograd, I.J; Riggs, A.C.; Coplen, T.B. 1998. The relative contributions of summer and cool- 4517 season precipitation to groundwater recharge, Spring Mountains, Nevada. Hydrogeology 4518 Journal. 6: 77–93. 4519 Woodward, A.; Schreiner, E.G.; Silsbee, D.G. 1995. Climate, geography, and tree 4520 establishment in subalpine meadows of the Olympic Mountains, Washington, USA. Arctic 4521 and Alpine Research. 27: 217–225. 4522 Zielinski, W.J., Truex, R.L., Schmidt, G.A. [et al.]. 2004. Home range characteristics of 4523 fishers in California. Journal of Mammalogy. 85: 649–657. 4524 Zielinski, W.J.; Duncan, N.P. 2004. Diets of sympatric populations of American martens 4525 (Martes americana) and fishers (Martes pennanti) in California. Journal of Mammalogy. 85: 4526 470–477.

DRAFT 105

Hudec et al. Chapter 6

4527 Chapter 6: Effects of Climate Change on Recreation in Southwest 4528 Washington 4529 4530 4531 Michael S. Hand, David L. Peterson, Nikola Smith, Becky P. Blanchard, Deb Schoenberg, and 4532 Robin Rose1 4533 4534 4535 Introduction 4536 4537 Public lands provide opportunities for outdoor recreation and connections to nature. Outdoor 4538 recreation is increasingly recognized as a source of wide-ranging benefits, from economic 4539 expenditures that support national industries and local gateway communities to personal and 4540 social benefits such as improved health and wellbeing, cultural and spiritual practices, and 4541 sustained family ties and traditions. Access to recreation opportunities is a key consideration that 4542 shapes where people live, work, and travel in the western United States, including in western 4543 Washington where national forests and national parks offer year-round opportunities for outdoor 4544 recreation. This is also true in Southwest Washington, where the most recent data (2011) show 4545 that Gifford Pinchot National Forest and Mount St. Helens National Volcanic Monument had 4546 over 1 million visits (USDA FS 2016). 4547 National forests provide recreation opportunities at sites that offer a wide variety of 4548 characteristics across all seasons of the year (fig. 6.1). Recreation on public lands in Southwest 4549 Washington is inseparable from ecosystems and natural features. Whether skiing, hiking, 4550 hunting, or camping, visiting developed sites or the backcountry, or simply driving through the 4551 mountains, natural and ecological conditions in large part determine the overall recreation 4552 experience. 4553 Climatic conditions and environmental characteristics that depend on climate are key 4554 factors that determine the availability of and demand for different recreation opportunities (Shaw 4555 and Loomis 2008). Changing climate conditions may alter the supply and demand of recreation 4556 opportunities, resulting in changes in the pattern of and benefits derived from recreation in the 4557 future. Climate change is projected to increase summer and warm-weather based recreation

1 Michael S. Hand is a research economist, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station, 800 East Beckwith, Missoula, Montana 59801; David L. Peterson is a senior research biological scientist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, 400 N 34th St., Suite 201, Seattle, WA 98103; Nikola Smith is an ecologist and ecosystem services specialist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Region, 1220 SW 3rd Avenue, Portland, OR 97204; Becky P. Blanchard is the wilderness, Wild and Scenic Rivers, and Congressionally Designated Areas program manager, U.S. Department of Agriculture, Forest Service, Pacific Northwest Region, 1220 SW 3rd Avenue, Portland, OR 97204; Deb Schoenberg is the recreation planner, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, 501 E 5th Street #404, Vancouver, WA 98661; and Robin Rose is the recreation program manager, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, 501 E 5th Street #404, Vancouver,DRAFT WA 98661. 106

Hudec et al. Chapter 6

4558 participation (Bowker et al. 2013), especially in locations where snow-based winter activities are 4559 currently prevalent (Loomis and Crespi 2004, Mendelsohn and Markowski 2004). 4560 Although broad trends in recreation participation under climate change scenarios may be 4561 borne out at the regional scale (chapter 2), climate change-induced changes to recreational use in 4562 Southwest Washington are difficult to predict due to the complex interactions between visitor 4563 behavior and the natural landscape. This chapter describes the broad categories of recreation 4564 activities that may be sensitive to climate-related changes and assesses the likely effects of 4565 projected climate-related changes on recreation participation in Southwest Washington. 4566 4567 4568 Relationships Between Climate Change and Recreation 4569 4570 The supply of and demand for recreation opportunities are sensitive to climate through (1) direct 4571 effects of changes in temperature and precipitation on the availability and quality of recreation 4572 sites, and (2) and indirect effects of climate on the characteristics and ecological condition of 4573 recreation sites (Hand et al. 2016, Loomis and Crespi 2004, Mendelsohn and Markowski 2004, 4574 Shaw and Loomis 2008) (fig. 6.2). Direct effects of changes in temperature and precipitation 4575 patterns are likely to affect most outdoor recreation activities in some way. Direct effects are 4576 important for skiing and other snow-based winter activities that depend on seasonal temperatures 4577 and the amount, timing, and phase of precipitation (Englin and Moeltner 2004, Irland et al. 2001, 4578 Stratus Consulting 2009, Wobus et al. 2017). Warm-weather activities are also sensitive to direct 4579 effects of climate change. The number of projected warm-weather days is positively associated 4580 with anticipated visitation for U.S. national parks (Albano et al. 2013, Fisichelli et al. 2015), an 4581 inference that is likely relevant on other public lands as well. Temperature and precipitation may 4582 also directly affect the comfort and enjoyment that participants derive from engaging in an 4583 activity on a given day (Mendelsohn and Markowski 2004). 4584 Indirect climate effects tend to be important for recreation activities that depend on 4585 additional ecosystem inputs, such as wildlife, vegetation, and surface water. Cold-water fishing 4586 is expected to decline in the future due to climate effects on temperature and stream flow that 4587 threaten cold-water fish species habitat (Jones et al. 2013) (chapter 3). Surface water area and 4588 stream flows are important for water-based recreation (e.g., boating), and forested area affects 4589 several outdoor activities (e.g., camping and hiking) (Loomis and Crespi 2004). Recreation visits 4590 to sites with highly valued natural characteristics, such as glaciers or popular wildlife species, 4591 may be reduced in some future climate scenarios if the quality of those characteristics is 4592 threatened (Scott et al. 2007). However, the desirability of recreation compared to alternative 4593 locations also affects visitor choices, a factor that must be considered in evaluating the effects of 4594 climate change on visitor behavior. The indirect climate effect on disturbances, and wildfire in 4595 particular (chapter 4), may also play a role in recreation behavior, although the effect may be 4596 heterogeneous and vary over time (Englin et al. 2001). 4597 4598 4599 Recreation Patterns in Southwest Washington 4600 4601 Recreation is an important component of public land management in Southwest Washington. 4602 Recreational resources are managed to connect people with natural resources and cultural 4603 heritage, andDRAFT to adapt to changing social needs and environmental conditions. Recreation 107

Hudec et al. Chapter 6

4604 managers on national forests, and to some extent on other federal lands, aim to provide diverse 4605 recreation opportunities that span the Recreation Opportunity Spectrum, from modern and 4606 developed to primitive and undeveloped (Clark and Stankey 1979) (box 6.1). For lands managed 4607 by the U.S. Forest Service, sustainable recreation serves as a guiding principle for planning and 4608 management purposes (USDA FS 2010, 2012b). Recreation is included among other major 4609 multiple uses of national forests, such as timber products and livestock grazing. Sustainable 4610 recreation on federal lands seeks to maintain and enhance the benefits that quality recreation 4611 opportunities provide across large landscapes in perpetuity (USDA FS 2010). 4612 People participate in a wide variety of outdoor recreation activities in Southwest 4613 Washington. The National Visitor Use Monitoring (NVUM) program, conducted by the U.S. 4614 Forest Service to monitor recreation visitation and activity on national forests, identifies 27 4615 different categories of recreation in which visitors participate. These include a wide variety of 4616 activities and ways that people enjoy and use national forests and other public lands. The most 4617 common recreational activities in Southwest Washington are related to viewing natural features, 4618 walking or hiking, and hunting (USDA FS 2016) (box 6.2). 4619 All outdoor recreation activities depend to some degree, directly or indirectly, on climate 4620 conditions or environmental conditions that are determined by climate. For example, wildlife 4621 viewing depends on habitat availability, which is determined by patterns of temperature, 4622 precipitation and disturbance. As climate change affects these factors, the quantity, timing, and 4623 duration of wildlife viewing opportunities could shift. 4624 To assess how recreation patterns may change in Southwest Washington, categories of 4625 outdoor recreation activities are identified that may be sensitive to climate changes. In this 4626 assessment, a recreation activity is sensitive to climate change if changes in climate, or 4627 environmental conditions that depend on climate, would result in a significant change in the 4628 demand for or supply of a recreation activity. Climate-sensitive recreation categories are 4629 addressed by season (warm-weather and cold-weather) as well as type (related to wildlife, 4630 gathering forest products, and water uses). 4631 4632 4633 Assessing Climate Change Effects on Recreation 4634 4635 This section provides an assessment of the likely effects of climate on climate-sensitive 4636 recreation activities in Southwest Washington. Two sources of information are used to develop 4637 assessments for each category of recreation activity. First, reviews of existing studies of climate 4638 change effects on recreation and studies of how recreation behavior responds to climate-sensitive 4639 ecological characteristics are used to draw inferences about likely changes for each activity 4640 category. Second, projections of ecological changes specific to Southwest Washington, as 4641 detailed in the other chapters contained in this publication, are paired with the recreation 4642 literature to link expected responses of recreation behavior to specific expected climate effects. 4643 Current conditions reflect wide variation in interannual and seasonal weather and 4644 ecological conditions. In essence, temperature, precipitation, water flows and levels, wildlife 4645 distributions, vegetative conditions, and wildfire activity exhibit wide ranges of variation. Thus, 4646 recreationists are already accustomed to making decisions with some degree of uncertainty about 4647 conditions at the time of participation. 4648 Recreation in Southwest Washington is affected by several existing challenges and 4649 stressors. DRAFTIncreasing population near urban areas, particularly those in close proximity to public 108

Hudec et al. Chapter 6

4650 lands, can strain visitor services and facilities; projected population increases in the future may 4651 exacerbate these effects (Bowker et al. 2012). Increased use from population growth can also 4652 reduce site quality because of congestion at the most popular sites (Yen and Adamowicz 1994). 4653 The physical condition of recreation sites and natural resources is constantly changing 4654 from human and natural forces. Recreation sites and physical assets need maintenance, and 4655 deferred or neglected maintenance may increase congestion at other sites that are less affected or 4656 increase hazards for visitors who continue to use degraded sites. Moreover, deferred or neglected 4657 maintenance can decrease user experience and cause unintended resource damage (e.g., adjacent 4658 aquatic resources). Unmanaged recreation can create hazards and contribute to natural resource 4659 degradation (USDA 2010). This stressor may interact with others, such as population growth and 4660 maintenance needs, if degraded site quality or congestion encourages users to engage in 4661 recreation that is not supported or appropriate at certain sites or at certain times of the year. 4662 Furthermore, natural hazards and disturbances may create additional challenges for the provision 4663 of recreation opportunities. For example, wildfire affects recreation demand (because of altered 4664 site quality and characteristics) and may also damage physical assets or exacerbate other safety 4665 hazards such as trees susceptible to falling near recreation areas and roads (hazard trees). 4666 The overall effect of climate change on recreation activity in Southwest Washington is 4667 likely to be (1) an increase in participation in warm-weather activities because of warmer 4668 temperatures and increased season length (Fisichelli et al. 2015), and (2) decreased winter 4669 activities that depend on snow because of decreased snowpack (Mendelsohn and Markowsi 4670 2004). However, these general inferences mask potential variation in the effects of climate on 4671 recreation between types of activities and geographic locations. 4672 4673 4674 Warm-Weather Activities 4675 4676 Warm-weather activities such as hiking, camping, and nature viewing are the most common 4677 recreation activities in Southwest Washington. Warm-weather recreation is sensitive to the 4678 seasonal duration of moderate weather conditions (Fisichelli et al. 2015), depending on the 4679 availability of snow- and ice-free trails and sites, and the timing and number of days with 4680 temperatures within a minimum and maximum comfortable range (which may vary with activity 4681 type and site). For example, the number of warm-weather days was a significant predictor of 4682 expected visitation behavior in Rocky Mountain National Park (Colorado) (Richardson and 4683 Loomis 2004), and minimum temperature was a strong predictor of monthly visitation patterns in 4684 Waterton Lakes National Park (Alberta, Canada) (Scott et al. 2007). Projections of lower 4685 precipitation during summer (chapter 2) may also contribute to conditions that are conducive to 4686 warm-weather activities. 4687 Overall demand for warm-weather activities is expected to increase because of the direct 4688 effect of a warmer climate on season length. For example, higher minimum temperatures are 4689 associated with increased number of hiking days (Bowker et al. 2012). Temperatures are 4690 predicted to increase 3-7 ºC across the region by the year 2100 (chapter 2), which is expected to 4691 result in earlier availability of snow- and ice-free sites and an increase in the number of warm- 4692 weather days in spring and autumn. This scenario of more snow-free warm-weather days in 4693 spring and autumn was borne out in Southwest Washington in 2014-2015, when snow-free 4694 access caused more widespread use of national forest lands at a time when field staff were not 4695 available DRAFTto manage recreation use. 109

Hudec et al. Chapter 6

4696 Extreme heat during summer months may shift demand to cooler weeks at the beginning 4697 or end of the warm-weather season, or shift demand to alternative sites that are less exposed to 4698 extreme temperatures. During periods of warm weather, Gifford Pinchot National Forest staff 4699 have noticed a greater concentration of water-related uses along river corridors and higher 4700 elevation areas that provide thermal relief. They have also observed increased rates of use in 4701 caves, such as Ape Cave, which is already a popular site that is sensitive to visitor impacts. 4702 Indirect effects of climate change on forested areas may have a negative impact on warm- 4703 weather recreation, primarily through wildfire effects, if site availability and quality are 4704 compromised. Wildfires have diverse and temporally nonlinear effects on recreation (Englin et 4705 al. 2001). The presence of recent wildfires has differential effects on the value of recreation 4706 experiences. Fire severity is significant: high-severity fires have been associated with decreased 4707 recreation visitation and decreased access, whereas low-intensity fires are associated with slight 4708 increases in visitation (Starbuck et al. 2006). 4709 Recent fires have been associated with initial losses of benefits for camping (Rausch et 4710 al. 2010) and backcountry recreation activities (Englin et al. 1996) that are attenuated over time. 4711 In 2015, Gifford Pinchot National Forest implemented fire restrictions prior to July 4 (no 4712 campfires outside of campgrounds, no motorized use on 400 km of trails) for the duration of the 4713 fire season because of ongoing fires and extreme fire hazard. Wildfires prompted additional 4714 closures of much of the Mt. Adams Wilderness and a reroute of the Pacific Crest Trail, despite 4715 an increase in demand by hikers. 4716 The effects of climate change on warm-weather recreation in Southwest Washington will 4717 depend on the condition of forest resources. Warmer weather and a longer snow-free season in 4718 the mountains will provide a longer season for activities such as hiking and camping. However, 4719 potential increases in wildfire activity may reduce demand for warm-weather activities in certain 4720 years because of degraded site desirability, impaired air quality from smoke, and limited site 4721 access. This was illustrated during the summers of 2015 and 2017, when widespread wildfire and 4722 smoke reduced access and the quality of recreation opportunities throughout much of 4723 Washington State. Southwest Washington is expected to have increased area burned by wildfire 4724 in future decades (chapter 4), which would have a negative impact on recreation visitation and 4725 benefits derived from recreation. 4726 More rain-on-snow events could damage campgrounds, roads and other infrastructure 4727 (fig. 6.3). The effects of rain-on-snow events was demonstrated by flooding that occurred in 4728 Southwest Washington in the winters of 2006-2007 and 2015-2016, which caused extensive 4729 damage in Gifford Pinchot National Forest and the surrounding area. 4730 Recreationists are also sensitive to site quality and characteristics, such as the presence 4731 and abundance of wildflowers, conditions of trails, and vegetation and cover (i.e., shade). The 4732 condition of unique features that are sensitive to climate change, such as snowpack and streams, 4733 may affect the desirability of certain sites (Scott et al. 2007). Many forested areas associated with 4734 warm-weather activities, such as camping, backpacking, hiking, and picnicking (Loomis and 4735 Crespi 2004), will be sensitive to a warmer climate (USDA 2012a, chapter 4) and may offer 4736 different characteristics in the future. 4737 Adaptive capacity among recreationists is high because of the large number of potential 4738 alternative sites. Some recreationists can alter the timing of visits and alter investments (e.g., 4739 appropriate gear), although some may be constrained by work schedules, family schedules, and 4740 finances. However, benefits derived from recreation may decrease even if substitute activities or 4741 sites are availableDRAFT (Loomis and Crespi 2004). For example, some alternative sites may involve 110

Hudec et al. Chapter 6

4742 greater difficulty of access due to remoteness or rugged terrain. Although the ability of 4743 recreationists to substitute sites and activities is well established, it is unclear how people 4744 substitute across time periods or between large geographic regions (e.g., choosing a site in 4745 Northeast Washington versus Southwest Washington) (Shaw and Loomis 2008). It is also 4746 unclear how much flexibility exists in scheduling outfitters and recreation concessionaires and if 4747 special-use permits can be modified to accommodate seasonal changes. 4748 4749 4750 Cold-Weather Activities 4751 4752 Snow-based recreation is highly sensitive to variations in temperature and the amount and timing 4753 of precipitation as snow. Seasonal patterns of temperature and snowfall determine the duration of 4754 time that a given site has sufficient snow for snow-based activities and is accessible (Scott et al. 4755 2008, Wobus et al. 2017). Lower temperatures and the presence of new snow are associated with 4756 increased demand for downhill skiing and snowboarding (Englin and Moeltner 2004), activities 4757 that occur in the backcountry and at developed sites including White Pass ski area on the north 4758 side of Gifford Pinchot National Forest (and partially on Okanogan-Wenatchee National Forest), 4759 and Mt. Hood Meadows ski area on Mt. Hood National Forest to the south. 4760 Climate change is expected to have a negative effect on snow-based winter activities, 4761 although a wide range of effects at local scales is possible because of variation in site location 4762 and elevation. Warmer winter temperatures are expected to reduce the proportion of precipitation 4763 as snow, even if the total amount of precipitation does not deviate significantly from historical 4764 norms (chapter 2). The rain-snow transition zone (where precipitation is more likely to be snow 4765 rather than rain for a given time of year) is expected to move to higher elevations, particularly in 4766 late autumn and early spring (Klos et al. 2014). This places lower elevation sites at risk of shorter 4767 or non-existent winter recreation seasons, potentially changing types and patterns of recreation. 4768 However, the highest elevation areas in the region are projected to remain snow-dominated for at 4769 least several decades in future climate scenarios (chapter 2). 4770 Studies of the ski industry in North America uniformly project negative effects of climate 4771 change (Scott and McBoyle 2007). Overall warming is expected to reduce season length and the 4772 likelihood of reliable winter recreation seasons. Climatological projections for Southwest 4773 Washington are consistent with studies of ski area vulnerability to climate change in other 4774 regions, where projected effects on skiing, snowboarding, and other snow-based recreation 4775 activities are negative (Dawson et al. 2009, Scott et al. 2008, Stratus Consulting 2009, Wobus et 4776 al. 2017). Low-elevation snow-based recreation areas will probably become less available for 4777 winter recreation, and those locations with adequate snow may face more recreation pressure 4778 (fig. 6.4). In Gifford Pinchot National Forest, some winter Sno-Park locations are at elevations 4779 where snow conditions are already less reliable for recreation than in the past. 4780 Snow-based recreationists have moderate capacity to adapt to changing conditions given 4781 the relatively large number of winter recreation sites in the region. A recent survey in Oregon 4782 showed that downhill skiers are willing to travel an average of 110 km to reach a ski area 4783 (Community Planning Workshop 2012), although this distance may be flexible if favorable snow 4784 conditions become more scarce. Although inter-regional substitution patterns for winter 4785 recreation, including increased expense and distance traveled, are poorly quantified (Shaw and 4786 Loomis 2008), changes in Southwest Washington sites relative to other regions may affect future 4787 visitation.DRAFT 111

Hudec et al. Chapter 6

4788 For activities in undeveloped or minimally developed sites (cross-country skiing, 4789 backcountry skiing, snowshoeing), recreationists may seek higher elevation sites with greater 4790 likelihoods of viable snow seasons (fig. 6.5). However, the infrastructure that currently provides 4791 access for skiing (e.g., trailheads, mostly at low to mid elevations) may not be well connected to 4792 persistent snow (mostly at high elevations) in the future. In January-February 2015, 4793 snowmobilers were observed in Mt. Adams Wilderness (Gifford Pinchot National Forest), where 4794 snow machines are prohibited. If this sort of activity, assumed to be a response to reduced 4795 snowpack at lower elevation, increases in the future, it will increase the likelihood of user 4796 conflicts. Lack of snow would also affect recreation events (snowmobile rallies, winter festivals), 4797 with possible implications for local economies that benefit from visitor spending. 4798 4799 4800 Wildlife Activities 4801 4802 Wildlife recreation activities involve terrestrial or aquatic animals as a primary component of the 4803 recreation experience. Wildlife recreation can involve consumptive (e.g., hunting) or non- 4804 consumptive (e.g., wildlife viewing, birding, catch-and-release fishing) activities. Distinct from 4805 other types of recreation, wildlife activities depend on the distribution, abundance, and quality of 4806 desired target species. These factors influence activity “catch rates,” that is, the likelihood of 4807 catching or seeing an individual of the target species. Sites with higher catch rates can reduce the 4808 costs associated with a wildlife activity (e.g., time and effort tracking targets) and enhance 4809 overall enjoyment of a recreation day for that activity (e.g., more views of highly valued 4810 species). 4811 Participation in wildlife activities is sensitive primarily to climate-related changes that 4812 affect expected catch rates. Catch rates are important determinants of site selection and trip 4813 frequency for hunting (Loomis 1995, Miller and Hay 1981), substitution among hunting sites 4814 (Yen and Adamowicz 1994), participation and site selection for fishing (Morey et al. 2002), and 4815 participation in non-consumptive wildlife recreation (Hay and McConnell 1979). Changes in 4816 habitat, food sources, or streamflows and water temperature (for aquatic species) may alter 4817 wildlife abundance and distribution, which in turn influences expected catch rates and wildlife 4818 recreation behavior. The general trend of declining fishing and hunting licenses in Washington 4819 over the past 20 years (USFWS 2017a,b) may reduce demand for animal harvest in Southwest 4820 Washington, regardless of any climate change effects. 4821 Wildlife activities may also be sensitive to other direct and indirect effects of climate 4822 change. The availability of highly valued targets affects some benefits derived from wildlife 4823 activities (e.g., bull trout [Salvelinus confluentus Suckley] for cold-water anglers) (Pitts et al. 4824 2012), as does species diversity for hunting (Milon and Clemmons 1991) and wildlife viewing 4825 (Hay and McConnell 1979). Temperature and precipitation are related to general trends in 4826 participation for multiple wildlife activities (Bowker et al. 2012, Mendelsohn and Markowski 4827 2004), although the exact relationship may be specific to activities and species. Some activities 4828 (e.g., big game hunting) may be enhanced by cold temperatures and snowfall at particular times 4829 to aid in field dressing, packing out harvested animals, and tracking. Other activities may be 4830 sensitive to direct climate effects similar to warm-weather activities, in which moderate 4831 temperatures and snow- and ice-free sites are desirable. 4832 Warming temperatures projected for Southwest Washington may increase participation in 4833 hunting, birding,DRAFT and viewing wildlife because of an increase in weather desirable for outdoor 112

Hudec et al. Chapter 6

4834 recreation (Bowker et al. 2012), although extreme heat would have a negative effect. Hunting 4835 that occurs during discrete seasons (e.g., elk [Cervus elaphus L.] and deer hunts) managed by 4836 state seasons may depend on weather conditions during a short period of time. The desirability of 4837 hunting during established seasons may decrease as warmer weather persists later into autumn 4838 and early winter and the likelihood of snow cover decreases, reducing harvest rates. In addition, 4839 the potential for conflicts with warm-weather recreation may increase, because hunting is not 4840 generally compatible with other forms of recreation. 4841 The effects of changes in habitat for target species are likely to be ambiguous because of 4842 complex relationships among species dynamics, vegetation, climate, and disturbances (e.g., 4843 wildfire). Vegetation productivity may decrease in the future, although this would probably have 4844 a neutral effect on game species populations, depending on vegetation composition and forage 4845 opportunities. Similarly, the effects of disturbances on harvest rates of target species are 4846 ambiguous because it is unknown exactly how habitat composition will change. 4847 Higher temperatures are expected to decrease populations of native cold-water fish 4848 species as climate refugia shrink to higher elevations (chapter 3). This change favors increased 4849 populations of fish species that can tolerate warmer temperatures. However, it is unclear whether 4850 shifting populations of species will affect catch rates, because relative abundance of fish may not 4851 change where warm-water species persist. 4852 Reduced snowpack and increased rain-on-snow would result in higher peak flows in 4853 winter and lower low flows in summer (chapter 2), creating stress for fish populations during 4854 different portions of their life histories (chapter 3). The largest patches of habitat for cold-water 4855 species will be at higher risk to shrink and fragment. Increased incidence and severity of wildfire 4856 may increase the likelihood of secondary erosion events that degrade waterways and fish habitat, 4857 and could affect infrastructure (e.g., docks, boat launches) used for fishing. These effects could 4858 degrade the quality of individual sites in a given year or decrease the desirability of angling as a 4859 recreation activity relative to other activities; however, some anglers will be able to shift 4860 activities to different sites and different target species. 4861 4862 4863 Gathering Forest Products 4864 4865 Gathering of forest products accounts for a relatively small portion of primary visitor activities in 4866 Southwest Washington and is relatively more common as a secondary activity. A small but avid 4867 population of enthusiasts for certain types of products supports the demand for gathering as a 4868 recreational activity. Small-scale commercial gathering likely competes with recreationists for 4869 popular and high-value products (e.g., huckleberries, some mushrooms), although resource 4870 availability may be sufficient to accommodate both types of gathering at current participation 4871 levels. In addition, members of federally recognized tribes are entitled through treaty rights, or 4872 through other agreements, to harvest various forest products and foods, often called “first foods,” 4873 from ceded territories and other traditional-use areas on federal lands. Gifford Pinchot National 4874 Forest works with several tribes to sustain and enhance huckleberry habitat (chapter 7). Seasons 4875 and cycles of harvesting may be altered by increased temperature and drought in the future. 4876 Forest product gathering is sensitive primarily to climatic and vegetative conditions that 4877 support the distribution and abundance of target species. Participation in forest product gathering 4878 is also akin to participation in warm-weather recreation activities, dependent on moderate 4879 temperaturesDRAFT and the accessibility of sites where products are typically found. Vegetative change 113

Hudec et al. Chapter 6

4880 caused by warming temperatures and variability in precipitation may alter the geographic 4881 distribution and productivity of target species (chapter 4). Increased incidence and severity of 4882 wildfires may eliminate current sources of forest products in some locations but may also 4883 encourage short- or medium-term productivity of other products (e.g., mushrooms). Higher 4884 elevation snow levels may allow more access to forest products collected outside of summer 4885 months, such as Christmas trees. 4886 Recreationists engaged in forest product gathering may have the ability to select different 4887 gathering sites as the distribution and abundance of target species changes, although alternate 4888 sites may have higher costs of gathering. Those who engage in gathering as a secondary activity 4889 may choose alternate activities to complement primary activities. 4890 4891 4892 Water-Based Activities, Not Including Fishing 4893 4894 Separate from angling, water-based activities comprise a small portion of primary recreation 4895 activity participation in Southwest Washington. Lakes and reservoirs provide opportunities for 4896 motorized and non-motorized boating and swimming, although boating is commonly paired with 4897 fishing. Upper reaches of streams and rivers, common on national forest land, are generally not 4898 desirable for boating and floating. Existing stressors include the occurrence of drought 4899 conditions that reduce water levels and site desirability in some years and disturbances that can 4900 alter water quality (e.g., erosion events following wildfires or damage from flooding). Water as a 4901 scenic resource, especially waterfalls, may also be disrupted. At least one popular waterfall 4902 (Curly Creek Falls) in Gifford Pinchot National Forest has been less prominent in late summer 4903 during the past decade. 4904 Warmer temperatures are expected to increase demand for water-based recreation as the 4905 viable season lengthens. Availability of suitable sites for water-based recreation is sensitive to 4906 reduced water levels caused by warming temperatures and decreased precipitation as snow. 4907 Lower surface-water area is associated with less participation in boating and swimming in 4908 particular (Bowker et al. 2012, Loomis and Crespi 2004, Mendelsohn and Markowski 2004), and 4909 streamflow magnitude is positively associated with number of days spent rafting, canoeing, and 4910 kayaking (Loomis and Crespi 2004). Demand for water-based recreation is also sensitive to 4911 temperature. Warmer temperatures are generally associated with higher participation in water- 4912 based activities (Loomis and Crespi 2004, Mendelsohn and Markowski 2004), although extreme 4913 heat may dampen participation for some activities (Bowker et al. 2012). However, the effects of 4914 climate change on water-based activities are expected to be small compared to the effects of 4915 changes in human demography and economic conditions. 4916 Increasing temperatures, reduced storage of water as snowpack, and increased variability 4917 of precipitation may lead to generally lower and more variable water levels in lakes and 4918 reservoirs on federal lands, which is associated with reduced site quality and suitability for 4919 certain activities. The susceptibility of lakes to reduced water quantity may differ depending on 4920 the reliability of the water source (e.g., springs versus streams). Warmer water is often associated 4921 with algal blooms in lakes, which reduce dissolved oxygen, decrease clarity, and may harm some 4922 aquatic species, humans, and pets. Algal blooms are already a management issue in some 4923 portions of the Pacific Northwest (Hand et al. in press), and impacts could increase as 4924 temperatures warm. Increased demand for surface water by downstream users may exacerbate 4925 low waterDRAFT levels in drought years. This may result in loss of developed or potable water systems 114

Hudec et al. Chapter 6

4926 in developed recreation sites and campgrounds. In addition, competition for water is expected to 4927 increase for other uses, such as hydroelectric power, which may lead to proposals for raising dam 4928 levels and altering water storage management. 4929 4930 4931 Summary 4932 4933 Several recreation activities are considered sensitive to changes to climate and ecosystem 4934 characteristics (table 6.1). However, recreation in Southwest Washington is diverse, and the 4935 effects of climate are likely to vary among different categories of activities and across 4936 geographic areas within the region. Participation in warm-weather recreation activities will likely 4937 increase (especially at higher elevations), primarily because higher temperatures would result in 4938 longer duration warm-weather periods and decreased snowpack, lengthening the duration of 4939 recreation site availability. In contrast, receding snow-dominated areas and shorter seasons will 4940 likely reduce opportunities (in terms of available days and sites) for winter recreation (especially 4941 at low to mid elevations). Less certainty exists about wildlife-based activities, forest product 4942 gathering, and water-based activities, because relationships between climate and these activities 4943 are not well understood, making it difficult to project climate-induced changes. 4944 Recreation demand is governed by several economic decisions with multiple interacting 4945 dependencies on climate. For example, on the demand side, decisions about whether to engage in 4946 winter recreation, which activity to participate in (e.g., downhill versus cross-country skiing), 4947 where to ski, how often to participate, and how long to stay for each trip depend to some degree 4948 on climate and ecological characteristics. On the supply side, site availability and quality depend 4949 on climate, but the effect may differ greatly from one location to another. Thus, climate effects 4950 on recreation depend on spatial and temporal relationships among sites, climate and ecological 4951 characteristics, and human decisions. Gifford Pinchot National Forest will need to adapt to 4952 changing visitor-use patterns to ensure sustainable recreation in Southwest Washington. Greater 4953 flexibility in hiring of seasonal employees, scheduling concessionaires, and providing access 4954 (roads, trails, campgrounds) will likely be needed to accommodate different and perhaps less 4955 predictable recreation demands (chapter 8). 4956 Uncertainty derives from unknown effects of climate on site quality and characteristics 4957 that are important for some recreation decisions (e.g., indirect effects of climate on vegetation, 4958 wildlife habitat, and species abundance and distribution). The exact effect of climate on target 4959 species or other quality characteristics may be spatially heterogeneous, yet these characteristics 4960 play a role in recreation decisions. 4961 Another source of uncertainty is how people will adapt to changes when making 4962 recreation decisions. Although substitution behavior over space and time is not well understood 4963 (Shaw and Loomis 2008), it may be important for Southwest Washington if some sites exhibit 4964 relatively little effect from climate change compared with sites in nearby regions. For example, 4965 winter recreation sites in Southwest Washington may experience shorter or lower-quality 4966 conditions in the future, but see increased demand if the quality of sites to the south becomes 4967 relatively worse. Many recreation activities that are popular in the region may have alternate 4968 sites, or timing of visits can be altered to respond to climate changes. Although the level of 4969 recreation participation may change little, alternative sites may not provide the same quality of 4970 experienceDRAFT (Loomis and Crespi 2004). If human population continues to increase in western 115

Hudec et al. Chapter 6

4971 Oregon and Washington, demand for recreation will probably increase even if the quality of 4972 some recreation opportunities declines. 4973 4974 4975 Literature Cited 4976 4977 Albano, C.M.; Angelo, C.L.; Strauch, R.L.; Thurman, L.L. 2013. Potential effects of 4978 warming climate on visitor use in three Alaskan national parks. Park Science. 30: 37–44. 4979 Bowker, J.W.; Askew, A.E.; Cordell, H.K. [et al.]. 2012. Outdoor recreation participation in 4980 the United States—projections to 2060: a technical document supporting the Forest Service 4981 2010 RPA assessment. Gen. Tech. Rep. GTR-SRS-160. Asheville, NC: U.S. Department of 4982 Agriculture, Forest Service, Southern Research Station. 42 p. 4983 Bowker, J.M.; Askew, A.E.; Poudyal, N.C. [et al.]. 2013. Climate change and outdoor 4984 recreation participation in the Southern United States. In: Vose, J.M.; Klepzig, K.D., eds. 4985 Climate change adaptation and mitigation management options: a guide for natural resource 4986 managers in Southern forest ecosystems. Boca Raton, FL: CRC Press. 32 p. 4987 Clark, R.N.; Stankey, G.H. 1979. The Recreation Opportunity Spectrum: a framework for 4988 planning, management, and research. Gen. Tech. Rep. PNW-98. Portland, OR: U.S. 4989 Department of Agriculture, Forest Service, Pacific Northwest Forest and Range Experiment 4990 Station. 4991 Community Planning Workshop. 2012. Oregon skier profile and economic impact analysis. 4992 Eugene, OR: University of Oregon. http://www.skioregon.org/wp-content/uploads/ski- 4993 oregon-econ-impact-final.pdf. (17 January 2017). 4994 Dawson, J.; Scott, D.; McBoyle, G. 2009. Climate change analogue analysis of ski tourism in 4995 the northeastern USA. Climate Research. 39: 1–9. 4996 Englin, J; Boxall, P.C.; Chakraborty, K; Watson, D.O. 1996. Valuing the impacts of forest 4997 fires on backcountry forest recreation. Forest Science. 42: 450–455. 4998 Englin, J.; Loomis, J.; González-Cabán, A. 2001. The dynamic path of recreational values 4999 following a forest fire: A comparative analysis of states in the Intermountain West. Canadian 5000 Journal of Forest Research. 31: 1837–1844. 5001 Englin, J.; Moeltner, K. 2004. The value of snowfall to skiers and boarders. Environmental and 5002 Resource Economics. 29: 123–136. 5003 Hand, M.L.; Lawson, M. 2018. Effects of climate change on recreation. In: Halofsky, J.E.; 5004 Peterson, D.L.; Dante-Wood, S.K. [et al.], eds. Climate change vulnerability and adaptation 5005 in the Northern Rockies. Gen. Tech. Rep. RMRS-GTR-374. Fort Collins, CO: U.S. 5006 Department of Agriculture, Forest Service, Rocky Mountain Research Station. 5007 Hand, M.S.; Peterson, D.L.; Blanchard, B.P. [et al.]. In press. Effects of climate change on 5008 recreation. In: Halofsky, J.E.; Peterson, D.L.; Ho, J.J., eds. Climate change vulnerability and 5009 adaptation in south central Oregon. Gen. Tech. Rep. PNW-GTR-xxx. Portland, OR: U.S. 5010 Department of Agriculture, Forest Service, Pacific Northwest Research Station. In press. 5011 Hay, M.J.; McConnell, K.E. 1979. An analysis of participation in nonconsumptive wildlife 5012 recreation. Land Economics. 55: 460–471. 5013 Irland, L.C.; Adams, D.; Alig, R. [et al.]. 2001. Assessing socioeconomic impacts of climate 5014 change on U.S. forests, wood-product markets, and forest recreation. BioScience. 51: 753– 5015 764. DRAFT 116

Hudec et al. Chapter 6

5016 Jones, R.; Travers, C.; Rodgers, C. [et al.]. 2013. Climate change impacts on freshwater 5017 recreational fishing in the United States. Mitigation and Adaptation Strategies for Global 5018 Change. 18: 731–758. 5019 Klos, P.Z.; Link, T.E.; Abatzoglou, J.T. 2014. Extent of the rain-snow transition zone in the 5020 western U.S. under historic and projected climate. Geophysical Research Letters. 41: 4560– 5021 4568. 5022 Loomis, J. 1995. Four models for determining environmental quality effects on recreational 5023 demand and regional economics. Ecological Economics. 12: 55–65. 5024 Loomis, J.; Crespi, J. 2004. Estimated effects of climate change on selected outdoor recreation 5025 activities in the United States. In: Mendelsohn, R.; Neumann, J., eds. The impact of climate 5026 change on the United States economy, Cambridge, United Kingdom: Cambridge University 5027 Press: 289–314. 5028 Mendelsohn, R.; Markowski, M. 2004. The impact of climate change on outdoor recreation. In: 5029 Mendelsohn, R.; Neumann, J., eds. The impact of climate change on the United States 5030 economy, Cambridge, United Kingdom: Cambridge University Press: 267–288. 5031 Miller, J.R.; Hay, M.J. 1981. Determinants of hunter participation: duck hunting in the 5032 Mississippi flyway. American Journal of Agricultural Economics. 63: 401–412. 5033 Milon, J.W.; Clemmons, R. 1991. Hunters’ demand for species variety. Land Economics. 67: 5034 401–412. 5035 Morey, E.R.; Breffle, W.S.; Rowe, R.D.; Waldman, D.M. 2002. Estimating recreational trout 5036 fishing damages in Montana’s Clark Fork river basin: summary of a natural resource damage 5037 assessment. Journal of Environmental Management. 66: 159–170. 5038 Pitts, H.M.; Thacher, J.A.; Champ, P.A.; Berrens, R.P. 2012. A hedonic price analysis of the 5039 outfitter market for trout fishing in the Rocky Mountain West. Human Dimensions of 5040 Wildlife. 17: 446–462. 5041 Rausch, M.; Boxall, P.C.; Verbyla, A.P. 2010. The development of fire-induced damage 5042 functions for forest recreation activity in Alberta, Canada. International Journal of Wildland 5043 Fire. 19: 63–74. 5044 Richardson, R.B.; Loomis, J.B. 2004. Adaptive recreation planning and climate change: A 5045 contingent visitation approach. Ecological Economics. 50: 83–99. 5046 Scott, D.; Dawson, J.; Jones, B. 2008. Climate change vulnerability of the U.S. Northeast 5047 winter recreation-tourism sector. Mitigation and Adaptation Strategies for Global Change. 5048 13: 577–596. 5049 Scott, D.; Jones, B.; Konopek, J. 2007. Implications of climate and environmental change for 5050 nature-based tourism in the Canadian Rocky Mountains: a case study of Waterton Lakes 5051 National Park. Tourism Management. 28: 570–579. 5052 Scott, D.; McBoyle, G. 2007. Climate change adaptation in the ski industry. Mitigation and 5053 Adaptation Strategies for Global Change. 12: 1411–1431. 5054 Shaw, D.; Loomis, J. 2008. Frameworks for analyzing the economic effects of climate change 5055 on outdoor recreation and selected estimates. Climate Research. 36: 259–269. 5056 Starbuck, C.M.; Berrens, R.P.; McKee, M. 2006. Simulating changes in forest recreation 5057 demand and associated economic impacts due to fire and fuels management activities. Forest 5058 Policy Economics. 8: 52–66. 5059 Stratus Consulting. 2009. Climate change in Park City: an assessment of climate, snowpack, 5060 and economicDRAFT impacts. Report prepared for The Park City Foundation. Washington, DC: 117

Hudec et al. Chapter 6

5061 Stratus Consulting, Inc. http://www.parkcitymountain.com/site/mountain- 5062 info/learn/environment. (17 January 2017). 5063 U.S. Department of Agriculture, Forest Service [USDA FS]. 1990. ROS primer and field 5064 Guide. Washington, DC: Recreation staff. 5065 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010. Connecting people with 5066 America’s great outdoors: a framework for sustainable recreation. 5067 http://www.fs.usda.gov/Internet/FSE_DOCUMENTS/stelprdb5346549.pdf. (17 January 5068 2017). 5069 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012a. Future of America’s 5070 forest and rangelands: Forest Service 2010 Resources Planning Act assessment. Gen. Tech. 5071 Rep. WO-87. Washington, DC: U.S. Department of Agriculture, Forest Service. 198 p. 5072 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012b. National Forest System 5073 land management planning, 36 CFR Part 219, RIN 0596-AD02. Federal Register. 77(68), 5074 21162–21276. 5075 U.S. Fish and Wildlife Servce [USFWS]. 2017a. Historical fishing license data. 5076 https://wsfrprograms.fws.gov/subpages/licenseinfo/fishing.htm. (20 August 2017). 5077 U.S. Fish and Wildlife Servce [USFWS]. 2017b. Historical hunting license data. 5078 https://wsfrprograms.fws.gov/subpages/licenseinfo/hunting.htm. (20 August 2017). 5079 Wobus, C.; Small, E.E.; Hosterman, H. [et al.]. 2017. Projected climate change impacts on 5080 skiing and snowmobiling: a case study of the United States. Global Environmental 5081 Change. 14: 1-14. 5082 Yen, S.T.; Adamowicz, W.L. 1994. Participation, trip frequency and site choice: a multinomial- 5083 Poisson hurdle model of recreation demand. Canadian Journal of Agricultural Economics. 5084 42: 65–76.

DRAFT 118

Hudec et al. Chapter 7

5085 Chapter 7: Climate Change and Ecosystem Services in Southwest 5086 Washington 5087 5088 5089 Nikola Smith, Alec Kretchun, Christopher J. Donnermeyer, Jessica L. Hudec, and Tracy L. 5090 Calizon1 5091 5092 5093 Introduction 5094 5095 Ecosystem services are the benefits people receive from nature. They are critical building blocks 5096 of human societies. A global analysis of human dependence on natural systems known as the 5097 Millennium Ecosystem Assessment (MEA) found that 60 percent of these goods and services are 5098 declining faster than they can recover (MEA 2005). This is partly due to the fact that 5099 relationships between ecological conditions and flows of benefits are poorly understood or 5100 inadequately considered in resource decisionmaking. The MEA drew attention to these critical 5101 goods and services by highlighting their importance in four primary categories: (1) provisioning 5102 services such as food, fiber, energy and water; (2) regulating services including erosion and 5103 flood control, water purification and temperature regulation; (3) cultural services such as 5104 spiritual connections with the land, history, heritage, and recreation; and (4) supporting services 5105 or the foundations of systems such as soil formation, nutrient cycling and pollination. 5106 The effects of climate change on ecological systems will alter the ability of those systems 5107 to provide goods and services over time. Differential effects on ecosystem components, 5108 individual species, and species interactions will have implications for water availability and 5109 quality, regulation of flows and flood prevention, pollinator-plant relationships, forest products, 5110 and other benefits (Montoya and Raffaelli 2010, Mooney et al. 2009). A greater incidence of 5111 extreme climatic and disturbance events could significantly alter the ability of systems to provide 5112 goods and services on which people rely. Understanding the biological underpinnings of 5113 ecosystem services can help reduce the negative effects of climate change, increase resilience, 5114 and facilitate adaptation over time (Seidl et al. 2016). 5115 Efforts to integrate ecosystem services into U.S. Forest Service policy and practice have 5116 increased over the last several years. In 2013, the Forest Service chartered the National 5117 Ecosystem Services Strategy Team (USDA FS 2013). Comprised of scientists and managers 5118 within the National Forest System, State and Private Forestry, and the Pacific Northwest

1 Nikola Smith is an ecologist and ecosystem services specialist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Region, 1220 SW 3rd Avenue, Portland, OR 97204; Alec Kretchun is an ecosystem services program associate, U.S. Department of Agriculture, Forest Service, Pacific Northwest Region, 1220 SW 3rd Avenue, Portland, OR 97204; Christopher J. Donnermeyer is the heritage program manager, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650; Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650; and Tracy L. Calizon is the community engagement staff officer, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, 501 E 5th Street #404, Vancouver, WA 98661DRAFT. 119

Hudec et al. Chapter 7

5119 Research Station, this group is tasked with finding opportunities to incorporate ecosystem 5120 services into Forest Service programs and operations. The team has the lead in responding to a 5121 Presidential Memorandum issued in October, 2015 instructing federal agencies to incorporate 5122 ecosystem services into decisionmaking and requiring each agency to formalize a plan for doing 5123 so. 5124 The Forest Service 2012 Planning Rule (36 CFR 219) requires national forests to take 5125 ecosystem services into consideration in revising national forest land management plans. This 5126 chapter highlights the high priority climate change considerations for ecosystem services that 5127 may be considered during forest planning. From an operational standpoint, climate change 5128 vulnerability assessments are intended to inform the plan revision process by analyzing potential 5129 climate change effects relevant to land management. By including ecosystem services in climate 5130 change vulnerability assessments, the information gathered can more easily be incorporated 5131 when plan revision begins. 5132 The ecosystem services included in this chapter were selected in consultation with staff 5133 from Gifford Pinchot National Forest (GPNF). Based on a qualitative literature analysis along 5134 with available data within the assessment area, this chapter focuses on a subset of services based 5135 on their importance in and around the Southwest Washington landscape in order to make 5136 meaningful conclusions about the effects of climate change on these services. This mirrors the 5137 criteria outlined in the 2012 Planning Rule directives, which advise managers to focus on key 5138 ecosystem services in forest plan revision that are (1) important outside the planning area and (2) 5139 can be affected by U.S. Forest Service decisionmaking. Ecosystem services covered in this 5140 chapter are representative of all four categories (provisioning, regulating, cultural, supporting), 5141 thus providing a broad perspective on resource benefits. 5142 5143 5144 Forest Products 5145 5146 One of the primary responsibilities of the Forest Service has been to ensure a sustainable supply 5147 of forest products. Gifford Pinchot National Forest provides several wood products including 5148 timber, biomass, post and poles, and many special forest products (berries, foliage, etc.). Overall, 5149 timber output in GPNF follows the same pattern as much of the federal land in the Pacific 5150 Northwest. Output peaked at about 600 million board-feet in 1970, and then began to decline 5151 until 1992 when the impacts of the Northwest Forest Plan and competitive timber trading and 5152 shifts in demand on the global market caused timber cut volumes in the GPNF to decrease 5153 significantly (fig. 7.1). Since 1992, annual output has averaged 22 million board-feet, with 2015 5154 above average at 39 million board-feet. Recent annual timber harvest has been increasing, and is 5155 almost exclusively from thinning operations occurring in second growth stands. 5156 Currently, forest products in Southwest Washington are important for both commercial 5157 and noncommercial uses. The number of permits issued for non-timber forest products document 5158 the variety of ways in which the forest in being utilized (table 7.1, fig. 7.2). Firewood cutting, 5159 Christmas tree cutting, and mushroom collecting are among the most popular activities. Unique 5160 to GPNF are a high number of permits sold for gathering beargrass (Xerophyllum tenax), a plant 5161 with rich natural and cultural histories in the West (Hummel et al. 2012) (table 7.1). Revenue 5162 data also offer a different perspective on the importance of the special forest products program 5163 (fig. 7.2). However, revenue, by itself, is not a reliable indicator of the popularity of a particular 5164 activity. ForDRAFT example, bough sales, harvested primarily under contract, and which generate 120

Hudec et al. Chapter 7

5165 significant revenue from and for the forest, are underrepresented in permit numbers. Personal 5166 mushroom gathering requires a permit, but the permit is free. In 2017, the GPNF began requiring 5167 a free, self-issued personal-use huckleberry picking permit to gather clearer data on this activity. 5168 Previously, no permit was required because the activity is so popular, permit issuance would 5169 have been administratively burdensome. 5170 From a broad geographic perspective, climate change is projected to affect timber and 5171 forest products through changes in vegetation structure and growth, as well as altered disturbance 5172 (chapter 4). Increased physiological stress associated with higher temperatures are expected to 5173 reduce growth of low-elevation forests (Restaino et al. 2016), and long-term droughts may 5174 increase mortality in the driest locations (Allen et al. 2010). Increased frequency or severity of 5175 drought-related disturbances, such as insect outbreaks and wildfire, are also expected to cause 5176 widespread mortality (Seidl et al. 2008). These phenomena can alter forest productivity, 5177 potentially reducing the amount of merchantable timber and other forest products. Conversely, 5178 increased carbon dioxide concentrations and longer growing seasons could increase productivity 5179 of some species, especially in the subalpine zone, although experimental results are equivocal 5180 (Kirilenko and Sedjo 2007). Climate change may also induce a gradual shift in the distribution 5181 and abundance of tree species currently used to produce timber (Gonzalez et al. 2010). 5182 Biophysical changes will have implications for local and global socioeconomic 5183 conditions as well, affecting industries and communities that are dependent on timber and non- 5184 timber forest harvests. Local changes in supply and demand will be affected by climate change 5185 and global market fluctuations. Increased supply associated with stimulated production could 5186 lower commodity prices (Kirilenko and Sedjo 2007). Demand for timber will likely continue to 5187 grow slowly, while demand for biofuels may grow as nearby local economies seek alternative 5188 sources of energy. 5189 Climate change may affect special forest products through access and availability. Each 5190 individual plant species that provides these products will respond uniquely to climate change, 5191 affecting the quantity, quality, and seasonality of goods listed above. The magnitude and pace of 5192 these changes is uncertain. Access to forest products may shift as infrastructure management 5193 adapts to climate change. Changes in recreation patterns and other human activities as a result of 5194 climate change adaptation may also affect access to forest products. User group conflicts, 5195 particularly in years of low production and high demand, may increase if yields are low for 5196 several years in a row. Furthermore, loss of access may affect different user groups 5197 disproportionately, while shifting recreational patterns caused by climate change may also affect 5198 special forest product gathering (chapter 6). This could mean increased gathering in the shoulder 5199 seasons, creating additional demand for access to some recreation areas earlier in spring and later 5200 in autumn. Road conditions and potential closures can also affect gathering of recreational and 5201 commercial special forest products. 5202 5203 5204 Carbon 5205 5206 Carbon sequestration refers to the uptake and long-term storage of atmospheric carbon by forests 5207 and grasslands in biomass and soils. It is a dynamic process and one that involves both carbon 5208 uptake (via photosynthesis) and carbon release (via decomposition and disturbance). Carbon 5209 sequestration is considered a regulating ecosystem service, because it helps to mitigate carbon 5210 dioxide levelsDRAFT in the atmosphere. In this way, carbon storage in forests is “...becoming more 121

Hudec et al. Chapter 7

5211 valuable as the impacts of greenhouse gas emissions are becoming more fully understood and 5212 experienced” (USDA FS 2015). 5213 Forests of North America are currently a net carbon sink, absorbing more carbon than 5214 they are releasing (Pan et al. 2011). The National Forest System accounts for about 20 percent of 5215 all forest land area in the United States and about 25 percent of all carbon stored, with a net 5216 increase in total stock over time (USDA FS 2015). Management activities (e.g., prescribed fire, 5217 fuel reductions, thinning) typically represent a short-term carbon loss through the removal or 5218 burning of biomass (Birdsey and Pan 2015, Nunery and Keeton 2010). However, these short- 5219 term losses may help modify long-term carbon emissions by regulating the release of carbon in 5220 periodic small pulses, rather than in large pulses that accompany crown fires (Restaino and 5221 Peterson 2013). 5222 In response to a growing need for guidance on carbon management and stewardship, the 5223 Forest Service created a set of “carbon principles” designed to help resource managers and 5224 planners address carbon stewardship (USDA FS 2015) (box 7.1). The second of these six 5225 principles states, “Recognize carbon as one of many ecosystem services. Carbon sequestration is 5226 one of many benefits provided by forests, grasslands, and forest products now and in the future. 5227 Carbon sequestration should be considered in context with other ecosystem services.” It would, 5228 of course, be inappropriate to focus solely on carbon storage in the absence of other resource 5229 considerations. Rather, national forests are directed to quantify the state of the carbon resource, 5230 and how carbon stewardship might be blended with other natural resource goals in planning and 5231 management. Carbon estimates are most useful at very large spatial scales; baseline carbon 5232 estimates at the national forest scale do not fully inform needs for project-specific applications, 5233 although assessment of carbon stocks at the national forest scale may guide project-specific and 5234 NEPA analysis. 5235 Climate Change Performance Scorecard Element 9 (carbon assessment and stewardship) 5236 and the 2012 Planning Rule require national forests to identify baseline carbon stocks and to 5237 consider this information in resource management and planning. The Forest Service uses a 5238 nationally consistent carbon assessment framework to calculate carbon stocks (Woodall et al. 5239 2013), and estimates of total ecosystem carbon storage and flux have been produced for all 5240 national forests in the United States, based on data from the Forest Inventory and Analysis (FIA) 5241 program (USDA FS 2015). Carbon stock estimates reflect the amount of carbon stored in all 5242 forms of biomass as well as soil. Carbon stocks in the most recent year of calculation (2013) in 5243 GPNF are just over 200 Tg and have steadily risen 5-10 Tg per year since 2005 (USDA FS 2015) 5244 (fig. 7.3). 5245 Total forest ecosystem carbon (in all seven pools) stored in the Pacific Northwest Region 5246 increased from 2005 to 2013, with 2304 Tg in 2005 and reaching 2370 Tg in 2013 (figure 7.3). 5247 Figure 7.3 displays these trends for each of the national forests and grassland between the years 5248 2005 and 2013, where the Willamette National Forest stored the largest amount of carbon in the 5249 region, approximately 243 Tg in 2005 and 248 Tg in 2013. During this period, the Fremont, 5250 Gifford Pinchot, Malheur, Mt. Baker-Snoqualmie, Mt. Hood, Okanogan, Olympic, Rogue River, 5251 Siuslaw, Umpqua, Wenatchee, Winema, and Colville National Forests generally increased in 5252 total forest ecosystem carbon, while the Deschutes, Wallowa-Whitman, Ochoco, Siskiyou, and 5253 Umatilla National Forests generally decreased. Total forest ecosystem carbon in the Columbia 5254 River Gorge NSA and Crooked River National Grassland stayed the same throughout this period. 5255 Trends in forest carbon stocks throughout the West will be affected by direct 5256 physiologicalDRAFT effects on trees (e.g., reduced/increased productivity), and indirect climate- 122

Hudec et al. Chapter 7

5257 mediated impacts (e.g., increased disturbances, shifts in species and age composition) (Vose et 5258 al. 2012). Increasing frequency of droughts, wildfires, and insect outbreaks may lead to increase 5259 carbon losses via tree mortality and biomass volatilization (Allen et al. 2010, Bentz et al. 2010). 5260 However, there is potential for increased growth rates in many tree species, as well as expansion 5261 of forest to higher altitudes currently limited by climatic factors, although the carbon 5262 implications of these potential changes are uncertain. (See chapter 4 for more information on 5263 climate change effects on forest productivity and disturbances.) 5264 On-site sequestration through tree growth and biomass accumulation is not the only way 5265 in which carbon is stored by national forests. Harvested wood products (HWP), such as lumber, 5266 panels, and paper, can account for a significant amount of off-site carbon storage, and estimates 5267 of this addition are important for both national and regional level accounting (Skog 2008). 5268 Estimates at the regional level are presented in figure 7.4 for harvested wood products still in use 5269 and solid waste disposal sites. Storage in HWP peaked at 143 Mg C in 1995, with total storage of 5270 131 Mg C in 2013, the most recent estimate available. 5271 5272 5273 Air Quality 5274 5275 The effects of climate change on air quality depend on both long-term climate trends and 5276 atmospheric chemistry (Bytnerowicz et al. 2007). Greenhouse gases, including carbon dioxide 5277 and methane, are themselves considered air pollutants by the U.S. Environmental Protection 5278 Agency (EPA) because of the risk they pose to public health and welfare (US EPA 2009). 5279 Forests contribute to good air quality by removing pollutants and particulate matter, and by 5280 mitigating climate change through carbon sequestration (Baro et al. 2014, Melillo et al. 2014, 5281 Nowak et al. 2006). Good air quality provided by forests also contributes to recreation, because 5282 people are drawn to forestlands to escape polluted air (among other things) in urban areas. 5283 Furthermore, clean air provides additional recreational benefits of good visibility and the 5284 maintenance of scenic viewsheds and night sky observations. 5285 Over the last 30 years, levels of many air pollutants, such as tropospheric ozone and 5286 nitrogen oxides, have decreased across the nation as a result of effective pollution control 5287 measures (Bytnerowicz et al. 2007). Higher temperatures and shifting global circulation patterns 5288 are expected to increase ozone and fine particulate matter (Fann et al. 2016), particularly in 5289 urban areas. Another significant impact to air quality, particularly in the western United States, 5290 will likely be increased smoke and particulate matter from larger and more frequent wildfires 5291 (Melillo et al. 2014). This finding is supported by the report “The impacts of climate change on 5292 human health,” which states with high confidence the likelihood of increasing wildfire impacts 5293 on human health (Fann et al. 2016). Increased fire-borne pollutants have implications for public 5294 health and safety, visibility, recreation, and environmental justice for communities in and around 5295 national forests (Fann et al. 2016). 5296 The most robust air quality information in Southwest Washington comes from 5297 monitoring programs in Indian Heaven, Goat Rocks, and Mt. Adams Wilderness Areas. Air 5298 quality monitoring directly measures ambient gases and particulates, and lake and stream water, 5299 and indirectly measures pollutant effects in lichens and fish. Indian Heaven and Goat Rocks 5300 Wilderness Areas are included in the EPA Western Lakes Survey, which established a baseline 5301 for water chemistry against which the health of lakes in the western United States can be 5302 comparedDRAFT (Landers et al. 1987). All three wilderness areas are included in the region-wide lichen 123

Hudec et al. Chapter 7

5303 monitoring project, which samples lichens for known air pollutants and analyzes shifts in lichen 5304 community composition (Glavich 2016). 5305 The Wilderness Air Quality Plan of the U.S. Forest Service Pacific Northwest Region 5306 (USDA FS 2012) is based on goals of the 10-Year Wilderness Stewardship Challenge. A priority 5307 sensitive receptor for each wilderness area is listed, along with a rating of 1 to 10 (most impaired 5308 to least impaired; data as of June 2012) of air quality based on various monitoring efforts. The 5309 priority sensitive receptors and overall scores for the three wilderness areas are: Indian Heaven 5310 (water: 8), Goat Rocks (lichens: 6), and Mt. Adams (lichens: 8). A separate air quality 5311 monitoring program focuses on Goat Rocks Wilderness (Horner and Peterson 1993). 5312 Characteristics in the wilderness areas listed as potentially affected by air quality are: Indian 5313 Heaven (flora, fauna, water), Goat Rocks (views, flora, fauna, water), and Mt. Adams (views, 5314 flora). 5315 All three wilderness areas are represented by Interagency Monitoring of Protected Visual 5316 Environments (IMPROVE) monitoring stations, which measure visibility and identify causal 5317 agents for visibility impairment. Figure 7.5 shows the number of summer days (May-September, 5318 2000-2013) on which visibility was impacted by smoke as recorded by the monitoring station at 5319 White Pass, Washington. 5320 The effects of increased wildfire are the biggest potential concern for air quality in 5321 Southwest Washington, because particulate matter from wildfire can reduce visibility, affect 5322 recreation patterns, and affect respiratory health in vulnerable populations. Located on the 5323 westside of the Cascade Range, Southwest Washington may be less affected than communities 5324 east of the crest, but smoke impacts on air quality are nonetheless expected to rise. In addition, 5325 tropospheric ozone may increase as a function of higher temperatures, regardless of air quality 5326 regulations. 5327 5328 5329 Cultural Services 5330 5331 Cultural ecosystem services include connections between people and the land that may be 5332 intangible, such as spiritual enrichment, heritage, identity, and aesthetic experiences. They also 5333 include practices such as harvesting of “first foods” by members of federally recognized tribes 5334 and rituals in sacred places. People often develop connections, or a sense of place, with specific 5335 features or landscapes. Memories, interactions, and history play a role in attachment of visitors 5336 and residents to the land (Eisenhauer et al. 2000, Kruger and Jakes 2003). The attraction of these 5337 places and experiences can influence where people live, work, and recreate (Smith et al. 2011). 5338 The effects of climate change on ecological structures, processes, and functions may 5339 influence culturally important natural resources, places, and traditions (Hess et al. 2008, Lynn et 5340 al. 2011). Some populations may be more affected by climate change than others because of 5341 geographic location, degree of association with climate-sensitive environments, and specific 5342 cultural, economic, or political characteristics (Lynn et al. 2011). Native American tribes may be 5343 particularly vulnerable to climate shifts because of their cultural connections with ecosystems 5344 and specific plant and animal species, as well as their use of resources for subsistence (Cordalis 5345 and Suagee 2008). 5346 The GPNF lies within the traditional homelands of the Sahaptin-, Salishan-, and Upper 5347 Chinookan-speaking groups. Modern traditional users of the forest include the Cowlitz, 5348 Nisqually,DRAFT Puyallup, Squaxin Island, Umatilla, Upper Chehalis, Wenatchee, and Yakama people. 124

Hudec et al. Chapter 7

5349 Over 250 locations of traditional uses in GPNF have been identified, most of which are camps, 5350 resource areas, and trails and places with mythical and spiritual significance (Hajda et al. 1995). 5351 Harvesting of first foods and other food resources represents a critical relationship 5352 between native communities and GPNF. First foods include salmon, berries (especially 5353 huckleberries [Vaccinium spp.]), roots (e.g., common camas [Camassia quamash {Pursh} 5354 Greene]), and large mammals; they were and still are an integral part of native culture and 5355 tradition. Ceremonies and harvesting practices are seasonal and climate dependent. Access to 5356 these foods may become less predictable if abundance and distribution of culturally important 5357 species shifts. Salmon are particularly iconic for spiritual and economic significance for many 5358 Pacific Northwest tribes. Climate change is expected to reduce populations of some salmon 5359 species, especially at lower elevations, because of increased stream temperatures and altered 5360 streamflows (chapter 3). 5361 Huckleberries are sacred to many tribes in the Pacific Northwest, which regard berry 5362 gathering as a critically important spiritual, cultural, and social activity. The Forest Service 5363 established a “handshake agreement” with the Yakama Nation in 1932, setting aside 1,100 ha of 5364 traditional huckleberry patches (the Sawtooth berryfields) in the Twin Buttes Recreation Area 5365 (Fisher 1997). Although huckleberries now cover only about one-third of this area, the 5366 agreement represents a commitment by GPNF to cultural uses of the land. The Burley Mountain- 5367 Pole Patch berryfields (3,100 ha) are another traditional cultural property in GPNF. Travel into 5368 the mountains for the annual gathering is important to the historical identity of the Cowlitz 5369 Indian Tribe and Yakama Nation, among others. Ongoing restoration projects in GPNF are 5370 helping to re-establish huckleberries in areas that have been traditionally used for harvest. 5371 Climate change may affect food harvesting traditions, which tend to be seasonal and 5372 cyclical or associated with specific locations. August is historically the month for peak 5373 huckleberry ripeness and collection, accompanied by social gatherings and trading. Berry 5374 collection usually occurs at elevations of 900-1,500 m (Hajda et al. 1995). Hunting and salmon 5375 fishing align with seasons, locations, and weather conditions, occurring at 300-600 m (Hajda et 5376 al. 1995), with camps typically in close proximity to streams, lakes, springs, and marshes. Shifts 5377 in hydrology and phenology may influence the characteristics of natural features and the timing 5378 and yield of traditional activities (CIER 2007, Lynn et al. 2013). 5379 Places of special designation such as National Monuments and Wild and Scenic Rivers 5380 are important ecological and cultural components of the landscape for several communities. 5381 GPNF has 373 km of designated and recommended Wild and Scenic Rivers and one National 5382 Volcanic Monument (Mount St. Helens). The Land and Resource Management Plan for GPNF 5383 notes a high demand for free-flowing river-related recreation, which could be affected by lower 5384 summer streamflows in a warmer climate (chapter 6). 5385 Mount St. Helens is a defining feature of the Cascade Range, contributing to aesthetics, 5386 sense of place, and history in Southwest Washington. Called Lawetlat’la by the Cowlitz, it is 5387 significant to the origin and establishment of the Cowlitz, Yakama, and other tribes (McClure et 5388 al. 2013, McClure and Reynolds 2015). Periodic eruptions of Lawetlat’la are perceived as an 5389 expression of natural inner social turmoil, reflecting the interconnectedness of social, natural, 5390 and supernatural realms (McClure et al. 2013), and supporting its listing in the National Register 5391 of Historic Places. If this area experiences increasing recreation, as visitors seek refuge from heat 5392 in a warmer climate (chapter 6), conflicts with traditional uses of the land could occur. 5393 A long history of harvest by Native Americans forms the basis of traditional ecological 5394 knowledge,DRAFT which may include specific pieces of information or knowledge systems that emerge 125

Hudec et al. Chapter 7

5395 from symbiotic relationships between people and places that are unique to each tribe (CTKW 5396 2014). Such knowledge has helped tribes respond adaptively to past climate stressors, including 5397 maintenance of sustainable harvests during periods of historic reductions in salmon populations 5398 and habitat quality (Lynn et al. 2013). 5399 5400 5401 Plant Biological Diversity and Invasive Species 5402 5403 The direct effects of a warmer climate and indirect effects of increased disturbances are expected 5404 to alter the abundance and distribution of plant species and communities (chapter 4), thus 5405 affecting biological diversity as well as availability of economically valuable species (box 7.2). 5406 Nonnative plant species, especially those known to be invasive, may be able to outcompete 5407 native species in drier soil conditions and in open habitats created by wildfire and other 5408 disturbances, potentially altering evapotranspiration rates, streamflow (Pejchar and Mooney 5409 2009), and overall plant productivity (Eviner et al. 2012). Invasives may reduce biodiversity, 5410 thus influencing genetic resources in existing plant communities (Charles and Dukes 2007) and 5411 potentially affecting adaptive capacity. 5412 Invasives may alter a variety of ecosystem functions, with implications for system 5413 resilience and ecosystem services. Nonnative pollinators may displace native species that are 5414 superior at pollination, facilitating range expansion in pollinator-limited invasives and distracting 5415 pollinators away from natives (Charles and Dukes 2007, Pejchar and Mooney 2009). Altered 5416 species assemblages may also affect fire regimes and carbon sequestration. The best example of 5417 this is the increasing prevalence of cheatgrass (Bromus tectorum L.) in Western shrublands and 5418 forests, providing fine fuels that encourage more frequent fire and discourage propagation of 5419 native species (Pimentel et al. 2001, 2004). Altered plant assemblages can also influence 5420 biogeochemical cycling, with implications for site productivity (Eviner et al. 2012, Pejchar and 5421 Mooney 2009). 5422 Some riparian zone invasives can have a negative effect on water regulation, potentially 5423 altering channel morphology and decreasing water holding capacity, thus increasing flood risk 5424 (Charles and Dukes 2007, Eviner et al. 2012, Pejchar and Mooney 2009). Water quality can also 5425 be compromised by erosion if the root structure of invasives decreases soil stability. Altered 5426 composition and function of wetland plant species can in some cases damage water filtration, 5427 storage, and flow regulation. For example, reed canary grass (Phalaris arundinacea L.) reduces 5428 water storage capacity in marsh systems and degrades habitat for the yellow rail (Coturnicops 5429 noveboracensis Gmelin), a migratory bird with a limited range in the western United States. 5430 The nonnative Japanese knotweed (Fallopia japonica [Houtt.] Ronse Decr.) reduces the 5431 quality of recreation and tourism by forming dense stands that crowd out native species, thereby 5432 impeding access and reducing habitat quality for both native plant species and wildlife habitat 5433 (Charles and Dukes 2007). Both aquatic and terrestrial plants can interfere with watercraft, 5434 reduce water quality, and reduce the abundance and diversity of fish and wildlife (Eiswerth et al. 5435 2005). The presence of invasives and altered disturbance patterns can also influence scenic views 5436 and aesthetics as well as cultural and spiritual experiences (Charles and Dukes 2007). 5437 Interactions among invasive species, ecological structure and function, and ecosystem 5438 services are complex and vary through space and time. Assessments of these relationships 5439 require understanding the species or assemblages that are key service providers or degraders, and 5440 how they DRAFTrespond to changing climatic conditions (Eviner et al. 2012). Site-specific knowledge 126

Hudec et al. Chapter 7

5441 can assist in understanding the vulnerability of systems to invasion and their subsequent ability 5442 to provide ecosystem services that are critical to human wellbeing. 5443 5444 5445 Literature Cited 5446 5447 Allen, C.D.; Macalady, A.K.; Chenchouni, H. [et al.]. 2010. A global overview of drought and 5448 heat-induced tree mortality reveals emerging climate change risks for forests. Forest Ecology 5449 and Management. 259: 660–684. 5450 Baro, F.; Chaparro, L; Gomez-Baggethun, E. [et al.]. 2014. Contribution of ecosystem 5451 services to air quality and climate change mitigation policies: the case of urban forests in 5452 Barcelona, Spain. Ambio. 43: 466–479. 5453 Bentz, B.J.; Régnière, J.; Fettig, C.J. [et al.]. 2010. Climate change and bark beetles of the 5454 western United States: direct and indirect effects. BioScience. 60: 602–613. 5455 Birdsey, R.A.; Pan, Y. 2015. Trends in management of the world’s forests and impacts on 5456 carbon stocks. Forest Ecology and Management. 355: 83–90. 5457 Butler, E.; Stockmann, K.; Anderson, N. [et al.]. 2014. Estimates of carbon stored in 5458 harvested wood products from the U.S. Forest Service Pacific Northwest Region, 1909-2012. 5459 Missoula, MT: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 5460 Station. 5461 Bytnerowicz, A.; Omasa, K.; Paoletti, E. 2007. Integrated effects of air pollution and climate 5462 change on forests: a northern hemisphere perspective. Environmental Pollution. 147: 438– 5463 445. 5464 Centre for Indigenous Environmental Resources [CIER]. 2007. Climate change impacts on 5465 abundance of traditional foods and medicine—effects on a First Nation and their capacity to 5466 adapt. http://www.tribesandclimatechange.org/docs/tribes_498.pdf. (1 July 2015). 5467 Charles, H.; Dukes, J.S. 2007. Impacts of invasive species on ecosystem services. Ecological 5468 Studies. 19: 217–237. 5469 Climate and Traditional Knowledges Workgroup [CTKW]. 2014. Guidelines for considering 5470 traditional knowledges in climate change initiatives. http://climatetkw.wordpress.com. (5 5471 August 2015). 5472 Cordalis, D.; Suagee, D.B. 2008. The effects of climate change on American Indian and Alaska 5473 Native tribes. Natural Resources and Environment. 22: 45–49. 5474 Eisenhauer, B.W.; Krannich, R.S.; Blahna, D.J. 2000. Attachments to special places on public 5475 lands: an analysis of activities, meanings and community connections. Society and Natural 5476 Resources. 13: 421–441. 5477 Eiswerth, M.E.; Darden, T.D.; Johnson, W.S. [et al.]. 2005. Input-output modeling, outdoor 5478 recreation, and the economic impacts of weeds. Weed Science. 53: 130–137. 5479 Eviner, V.T.; Garback, K.; Baty, J.H.; Hoskinson, S.A. 2012. Measuring the effects of 5480 invasive plants on ecosystem services: challenges and prospects. Invasive Plants Science and 5481 Management. 5: 125–136. 5482 Fann, N.T., Brennan, P. Dolwick, J.L. [et al.]. 2016. Air quality impacts. In: Crimmins, A.; 5483 Balbus, J.; Gamble, J.L. [et al.], eds., The impacts of climate change on human health in the 5484 United States: a scientific assessment. Washington, DC: U.S. Global Change Research 5485 Program:DRAFT 69–98. 127

Hudec et al. Chapter 7

5486 Fisher, A.H. 1997. The 1932 handshake agreement: Yakama Indian treaty rights and Forest 5487 Service policy in the Pacific Northwest. Western Historical Quarterly (summer): 187–217. 5488 Glavich, D. 2016. Climate change monitoring with lichens on the west slope of the Cascade 5489 Range from southwest Washington to northwest Oregon. Portland, OR: U.S. Department of 5490 Agriculture, Forest Service, Pacific Northwest Region. 5491 Gonzalez, P.; Neilson, R.P.; Lenihan, J.M.; Drapek, R.J. 2010. Global patterns in the 5492 vulnerability of ecosystems to vegetation shifts due to climate change. Global Ecology and 5493 Biogeography. 19: 755–768. 5494 Hajda, Y.; Ellis, D.V.; Fagan, J.L. [et al.]. 1995. Ethnographic sites of the Gifford Pinchot 5495 National Forest, Washington, volume I. Unpublished report. Portland, OR: Archaeological 5496 Investigations Northwest. 5497 Hellmann, J.J.; Byers, J.E.; Bierwagen, B.G.; Dukes, J.S. 2008. Five potential consequences 5498 of climate change for invasive species. Conservation Biology. 22: 534–543. 5499 Hess, J.J.; Malilay, J.N.; Parkinson, A.J. 2008. Climate change: the importance of place. 5500 American Journal of Preventive Medicine. 35: 469–478. 5501 Horner, D.; Peterson, D.L. 1993. Goat Rocks Wilderness air quality monitoring plan. 5502 Packwood, WA: U.S. Department of Agriculture, Forest Service, Gifford Pinchot National 5503 Forest. 5504 Hummel, S.; Foltz-Jordan, S.; Polasky, S. 2012. Natural and cultural history of beargrass 5505 (Xerophyllum tenax). Gen. Tech. Rep. PNW-GTR-864. Portland, OR: U.S Department of 5506 Agriculture, Forest Service, Pacific Northwest Research Station. 5507 Kirilenko, A.P.; Sedjo, R.A. 2007. Climate change impacts on forestry. Proceedings of the 5508 National Academy of Sciences, USA. 104: 19697–19702. 5509 Kruger, L.E.; Jakes, P.J. 2003. The importance of place: advances in science and application. 5510 Forest Science. 49: 819–821. 5511 Landers, D.H.; Eilers, J.M.; Brakke, D.F. [et al.]. 1987. Western Lake Survey phase I: 5512 characteristics of lakes in the western United States, volume I: population descriptions and 5513 physico-chemical relationships. EPA Rep. 600/3-86/054a. Washington, DC: U.S. 5514 Environmental Protection Agency. 5515 Lynn, K.; Daigle, J.; Hoffman, J. [et al.]. 2013. The impacts of climate change on tribal 5516 traditional foods. Climatic Change. 120: 545–556. 5517 Lynn, K.; MacKendrick, K.; Donoghue, E.M. 2011. Social vulnerability and climate change: 5518 synthesis of literature. Gen. Tech. Rep. PNW-GTR-838. Portland, OR: U.S. Department of 5519 Agriculture, Forest Service, Pacific Northwest Research Station. 5520 McClure, R.; Reynolds, N. 2015. Making the list: Mount St. Helens as a traditional cultural 5521 property, a case study in tribal/government Cooperation. Journal of Northwest Anthropology. 5522 49: 117–142. 5523 McClure, R.; Hand, J.; Burke, A. 2013. National Register of Historic Places. Lawetlat’la / 5524 Mount St. Helens. National Park Service Form 10-900. Office of Management and Budget 5525 No. 1024-0018. https://www.nps.gov/nr/feature/places/pdfs/13000748.pdf. (8 February 5526 2018). 5527 Melillo, J.M.; Richmond, T.T.; Yohe, G. 2014. Climate change impacts in the United 5528 States. Third National Climate Assessment. Washington, DC: U.S. Global Change Research 5529 Program. 5530 Millennium Ecosystem Assessment [MEA]. 2005. Ecosystems and human well-being: 5531 biodiversityDRAFT synthesis. Washington, DC: World Resources Institute. 128

Hudec et al. Chapter 7

5532 Montoya, J.M.; Raffaeli, D. 2010. Climate change, biotic interactions and ecosystem services. 5533 Philosophical Transactions of the Royal Society of London B, Biological Sciences. 365: 5534 2013–2018. 5535 Mooney H.; Larigauderie A.; Cesario, M. [et al.]. 2009. Biodiversity, climate change, and 5536 ecosystem services. Current Opinions in Environmental Sustainability. 1: 46–54. 5537 Nowak, D.J.; Crane, D.E.; Stevens, J.C. 2006. Air pollution removal by urban trees and shrubs 5538 in the United States. Urban Forestry and Urban Greening. 4: 115–123. 5539 Nunery, J.S.; Keeton W.S. 2010. Forest carbon storage in the northeastern United States: net 5540 effects of harvesting frequency, post-harvest retention, and wood products. Forest Ecology 5541 and Management. 259: 1363–1375. 5542 Pan, Y. Birdsey, R.A., Fang. J. [et al.]. 2011. A large and persistent carbon sink in the world’s 5543 forests. Science. 333: 988–993. 5544 Pejchar, L.; Mooney, H.A. 2009. Invasive species, ecosystem services and human well-being. 5545 Trends in Ecology and Evolution. 24: 497–504. 5546 Pimentel, D.; McNair, S.; Wightman, J.J. [et al.]. 2001. Economic and environmental threats 5547 of alien plant, animal, and microbe invasions. Agriculture, Ecosystems and Environment. 84: 5548 1–20. 5549 Pimentel, D.; Zuniga, R.; Morrison, D. 2004. Update on the environmental and economic costs 5550 associated with alien-invasive species in the United States. Ecological Economics. 52: 273– 5551 288. 5552 Pyke, C.R.; Thomas, R.; Porter, R.D. [et al.]. 2008. Current practices and future opportunities 5553 for policy on climate change and invasive species. Conservation Biology. 22: 585–592. 5554 Restaino, C.M.; Peterson, D.L.; Littell, J.S. 2016. Increased water deficit decreases Douglas- 5555 fir growth throughout western US forests. Proceedings of the National Academy of Sciences, 5556 USA. 113: 9557–9562. 5557 Restaino, J.C.; Peterson, D.L. 2013. Wildfire and fuel treatment effects on forest carbon 5558 dynamics in the western United States. Forest Ecology and Management. 303: 46–60. 5559 Seidl, R.; Rammer, W.; Jager, D.; Lexer, M.J. 2008. Impact of bark beetle (Ips typographus 5560 L.) disturbance on timber production and carbon sequestration in different management 5561 strategies under climate change. Forest Ecology and Management. 256: 209–220. 5562 Seidl, R.; Spies, T.A.; Peterson, D.L. [et al.]. 2016. Searching for resilience: addressing the 5563 impacts of changing disturbance regimes on forest ecosystem services. Journal of Applied 5564 Ecology. 53: 120–129. 5565 Skog, K.E. 2008. Sequestration of carbon in harvested wood products for the United States. 5566 Forest Products Journal. 58: 56–72. 5567 Smith, N.; Deal, R.; Kline, J. [et al.]. 2011. Ecosystem services as a framework for forest 5568 stewardship: Deschutes National Forest overview. Gen. Tech. Rep. PNW-GTR-852. 5569 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 5570 Station. 5571 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012. Pacific Northwest Region 5572 wilderness air quality plan. 5573 http://www.fs.fed.us/air/documents/R6%20Wilderness_AQ_Plan_062812- 5574 508remediation.pdf. (25 January 2017). 5575 U.S. Department of Agriculture, Forest Service [USDA FS]. 2013. Forest Service national 5576 strategicDRAFT framework for invasive species management. FS-1017. Washington, DC. 129

Hudec et al. Chapter 7

5577 U.S. Department of Agriculture, Forest Service [USDA FS]. 2015. Baseline estimates of 5578 carbon stocks in forests and harvested wood products for the National Forest System units, 5579 Pacific Northwest Region. 5580 http://www.fs.fed.us/climatechange/documents/PacificNorthwestRegionCarbonAssessment.p 5581 df. (15 October 2016). 5582 U.S. Environmental Protection Agency [US EPA]. 2009. Endangerment and cause or 5583 contribute findings for greenhouse gases under section 202(a) of the Clean Air Act; Final 5584 Rule, 74 FR 66496 (December 15, 2009). https://www.gpo.gov/fdsys/pkg/FR-2009-12- 5585 15/pdf/E9-29537.pdf. (13 November, 2017). 5586 Vose, J.M; Peterson, D.L.; Patel-Weynand, T., eds. 2012. Effects of climatic variability and 5587 change on forest ecosystems: a comprehensive science synthesis for the U.S. forest sector. 5588 Gen. Tech. Rep. PNW-GTR-870. Portland, OR: U.S. Department of Agriculture, Forest 5589 Service, Pacific Northwest Research Station. 5590 Woodall, C.; Smith, J.; Nichols, M. 2013. Data sources and estimation/modeling procedures 5591 for the National Forest System carbon stock and stock change estimates derived from the US 5592 National Greenhouse Gas Inventory. Washington, DC: U.S. Department of Agriculture, 5593 Forest Service. 5594 https://www.fs.fed.us/climatechange/documents/NFSCarbonMethodology.pdf. (25 January 5595 2017).

DRAFT 130

Hudec et al. Chapter 8

5596 Chapter 8: Adapting to the Effects of Climate Change in Southwest 5597 Washington 5598 5599 5600 Jessica E. Halofsky and Jessica L. Hudec1 5601 5602 5603 Introduction 5604 5605 Climate change adaptation can be defined as, “adjustment in natural or human systems in 5606 response to actual or expected climatic stimuli or their effects” (McCarthy et al. 2001). 5607 Adaptation can help to help reduce the negative effects of climate change and transition 5608 organisms and ecosystems to new conditions. Incorporating climate change into daily action and 5609 planning is a new frontier for natural resources managers. However, Federal land management 5610 agencies in the United States are required to evaluate the potential risks associated with climate 5611 change and develop strategies to minimize climate change effects on their operations and 5612 mission (see chapter 1). This chapter details actionable adaptation strategies and tactics 5613 developed for Federal lands in southwestern Washington. 5614 The Southwest Washington Adaptation Partnership (SWAP) conducted a science-based 5615 climate change vulnerability assessment of natural resources in southwestern Washington and 5616 used the vulnerability assessment to develop adaptation options for natural resource management 5617 on federal lands in the region. The SWAP adaptation planning process included key steps 5618 outlined in Peterson et al. (2011b), including: (1) education on basic climate change science, 5619 integrated with knowledge of local resource conditions and issues (review); (2) evaluation of the 5620 sensitivity of specific natural resources to climate change (rank); (3) development and 5621 implementation of adaptation strategies and tactics (resolve). An initial meeting with leadership 5622 and resource managers from Gifford Pinchot National Forest and the Washington Department of 5623 Natural Resources involved review of basic climate change information in a local context. That 5624 meeting was followed by a vulnerability assessment process that evaluated potential effects of 5625 climate change on fish and aquatic habitat (chapter 3), vegetation (chapter 4), special habitats 5626 (chapter 5), recreation (chapter 6), and ecosystem services (chapter 7). 5627 These assessments set the stage for hands-on development of adaptation options by 5628 resource managers, partners, and interested publics in a workshop setting. Participants included 5629 resource managers from U.S. Forest Service, Washington Department of Natural Resources, 5630 Washington Department of Fish and Wildlife, Yakama Nation; partners from forest collaborative 5631 groups, private industry, Pinchot Institute, Cascade Forest Conservancy, Cowlitz Tribe, and 5632 University of Washington; and local community members. Workshop participants engaged in 5633 facilitated discussion and completed worksheets, adapted from Swanston and Janowiak (2012), 5634 identifying key climate change vulnerabilities and related adaptation strategies (overarching 5635 approaches for resource planning and management) and tactics (on-the-ground management 5636 actions). Participants were encouraged to use the Climate Change Adaptation Library (Halofsky

1 Jessica E. Halofsky is a research ecologist, University of Washington, College of the Environment, School of Environmental and Forest Sciences, Box 352100, Seattle, WA 98195- 2100; Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot NationalDRAFT Forest, Mt. Adams Ranger District, 2455 Hwy 141, Trout Lake, WA 98650. 131

Hudec et al. Chapter 8

5637 et al. 2018; http://adaptationpartners.org/library.php) for ideas on adaptation strategies and 5638 tactics. They also identified locations where tactics could be applied and situations that provide 5639 specific opportunities for implementation of tactics, where applicable. This chapter describes 5640 high-priority adaptation strategies and tactics developed in the workshop for each of the five 5641 resource areas covered in the vulnerability assessment. Chapter 9 describes next steps for 5642 implementation and monitoring. 5643 5644 5645 Adapting Aquatic Habitat Management to Climate Change in Southwest 5646 Washington 5647 5648 Several comprehensive reviews have documented strategies for increasing fish population and 5649 aquatic habitat resilience to changing climate in streams of the western United States (e.g., ISAB 5650 2007, Luce et al. 2012, Mantua and Raymond 2014, Rieman and Isaak 2010). Resource 5651 managers in other parts of the Pacific Northwest previously used this existing information as a 5652 basis for developing adaptation strategies and tactics for fish, which are synthesized in the 5653 Climate Change Adaptation Library. Adaptation workshop participants in southwestern 5654 Washington identified strategies and tactics from the Adaptation Library that were relevant and 5655 appropriate for southwest Washington and added additional tactics (table 8.1). 5656 With increasing winter air temperatures, more precipitation will fall as rain rather than 5657 snow, and snowpacks will be reduced (Hamlet and Lettenmaier 2007, Mote et al. 2003). These 5658 changes will result in higher winter peak streamflows, and extreme flows will be more frequent 5659 than they are now, causing considerable stress for some fish species (Hamlet and Lettenmaier 5660 2007, Mote et al. 2003). In response to increased peak streamflows in winter and early spring, 5661 general adaptation strategies include increasing spawning habitat resilience by restoring stream 5662 structure and processes and by reducing threats from roads and infrastructure, particularly in 5663 floodplains (table 8.1). Restoring stream and floodplain complexity and function can help to 5664 reduce impacts of high flows (Peterson and Halofsky 2018). For example, logjams can help to 5665 slow water velocity when flows are high, and large wood can be beneficial for salmonid parr 5666 winter survival (Luce et al. 2012). Reducing road density near streams and disconnecting roads 5667 from streams (i.e., preventing sediment delivery into streams by engineering roads and drainage 5668 structures to disperse runoff into areas with stable forest floors) can help to reduce negative 5669 impacts of roads on floodplains and aquatic habitat (Luce et al. 2012). Increasing culvert 5670 capacity to accommodate higher future flows and removing or modifying other barriers to fish 5671 passage could also help to increase access to fish spawning habitat (Mantua et al. 2011). Road 5672 surveys can be used to maintain up-to-date information on culvert function and erosion. 5673 To minimize the negative impacts of lower summer streamflows on habitat quality 5674 (chapter 3), a key strategy is to decrease fragmentation of the stream network so fish can access 5675 suitable habitat (Isaak et al. 2012, Luce et al. 2012) (table 8.1). For example, managers may want 5676 to identify road-stream crossings that impede fish movement and prioritize culvert replacement 5677 in those locations (Mantua et al. 2011). Upland and riparian vegetation can be managed to 5678 promote retention of water (Mantua et al. 2010, Peterson and Halofsky 2018). Trees can be 5679 maintained and planted in riparian areas to increase shade over streams and keep stream 5680 temperatures down (Luce et al. 2012). Water conservation measures, including increasing 5681 efficiency of irrigation, help to protect in-stream flows (Peterson and Halofsky 2018). Other 5682 measures DRAFTto increase habitat resilience to lower summer streamflows include restoring American 132

Hudec et al. Chapter 8

5683 beaver (Castor canadensis Kuhl) populations (Pollock et al. 2014, 2015), increasing off-channel 5684 habitat and protecting refugia in side channels (Mantua et al. 2010), and protecting hyporrheic 5685 flows and wetland-fed streams that maintain higher summer flows (Muir et al. 2018). 5686 Lower summer streamflows and higher temperatures with climate change will also lead 5687 to increased stream temperatures, which will increase thermal stress for cold-water fish species 5688 (chapter 3). Restoring and maintaining habitat quality and protecting cold water refugia will help 5689 to mitigate effects of increased stream temperatures (Isaak et al. 2012) (table 8.1). Warmer 5690 stream temperatures may lead to increased incidence of disease and parasites in native fish 5691 species (ISAB 2007). To limit spread of diseases and parasites in fish populations, options 5692 include increasing public education to eliminate disease vectors, direct treatment or removal of 5693 infected fish, and considering changes in hatchery practices. 5694 Increased precipitation, precipitation intensity and peak flows in winter will increase 5695 landslide risk (Goode et al. 2012), and increased landslides may lead to sedimentation in streams 5696 (Luce et al. 2012). Post-fire flooding and landslide events, which may increase with greater area 5697 burned, also lead to increased stream sedimentation (Goode et al. 2012). These increases may 5698 exceed the capacity of the streams to process (through transport and deposition) sediments and 5699 degrade aquatic habitats. To maintain stream sedimentation associated with these disturbance 5700 events within a natural range, managers can work to reduce sediment generated by roads. For 5701 example, culverts can be resized to accommodate increasing peak flows, and roads and other 5702 infrastructure likely to be affected by flooding and landslides can be modified or relocated 5703 (Mantua et al. 2011). Decommissioning roads can be considered for high-risk locations (e.g., 5704 floodplains). 5705 To reduce post-fire stream sedimentation, forest thinning and prescribed fire can be used 5706 proactively to reduce fire severity and extent in dry forests (Halofsky et al. 2018), where 5707 ecologically appropriate and resource needs dictate. After fire events, vegetation restoration can 5708 be conducted in highly-impacted riparian areas and adjacent uplands to prevent erosion (Luce et 5709 al. 2012). Other currently used post-fire actions that could facilitate post-fire recovery in aquatic 5710 systems include creating barriers to ash and sediment around key stream habitat areas and 5711 contour falling trees to capture sediment and ash in steep drainages. 5712 Finally, long-term monitoring is critical to evaluate the effectiveness of climate-informed 5713 resource management (Young et al. 2017). More and higher quality data are needed for 5714 streamflow (more sites), stream temperature (annual data from sensors maintained over many 5715 years), fish distributions, and road and culvert conditions. These data will improve status-and- 5716 trend descriptions and contribute to models that more accurately project responses to climate 5717 change and land management activities (Peterson and Halofsky 2018). 5718 5719 5720 Adapting Forest Vegetation Management to Climate Change in Southwest 5721 Washington 5722 5723 Disturbance is likely to be the major catalyst for ecosystem change in a warming climate (Millar 5724 and Stephenson 2015). Area burned may significantly increase in Southwest Washington (Littell 5725 et al. 2010), particularly in drier forest types such as in the grand fir (Abies grandis [Douglas ex 5726 D. Don] Lindl.) and Douglas-fir (Pseudotsuga menziesii [Mirb.] Franco) zones (chapter 4). 5727 Incorporating increased fire and other effects of climate change into management plans can help 5728 managersDRAFT to better prepare for change (Millar et al. 2007) (table 8.2). In dry forest types, active 133

Hudec et al. Chapter 8

5729 fuels management programs can help to promote desired effects of fire on ecosystems while 5730 reducing unwanted effects (e.g., damage to infrastructure or highly-valued habitat). Fuels 5731 management could also help to decrease severity of fire and transition ecosystems to new, more 5732 frequent fire regimes (table 8.2). Fuels management practices could include thinning with 5733 hazardous fuel treatment, prescribed fire, and implementing fuelbreaks in strategic locations 5734 (Peterson et al. 2011a). Decreasing forest density and increasing structural heterogeneity through 5735 thinning, prescribed fire, and managed wildfire has the added benefit of increasing forest 5736 resilience to drought and insect outbreaks (Sohn et al. 2016). 5737 With increased area burned, the area in early-seral forest condition will likely increase, 5738 and the area of late-seral forest will likely decrease (Chmura et al. 2011, chapters 4 and 5). Thus, 5739 it may be necessary to actively manage, protect, and develop late-seral forest structure to 5740 maintain desired levels of late-seral habitat on the landscape (table 8.2). In forest types with 5741 infrequent, stand-replacing fire regimes, such as western hemlock (Tsuga heterophylla [Raf.] 5742 Sarg.) and Pacific silver fir (Abies amabilis Douglas ex J. Forbes) forests, continued fire 5743 suppression could help protect late-seral forest structure. In early- to mid-seral forest, active 5744 management could be used to facilitate development of late-seral structure (table 8.2). 5745 Climate change may alter conditions such that genetic material is no longer optimal for a 5746 given site. Thus, managers may need to reconsider the genetic material that is used for planting 5747 and consider material from multiple seed zones (table 8.2). With increased area burned, there 5748 will be a greater need for seed sources and propagated plants for post-fire planting. Managers 5749 could collect seed from trees that have desirable traits (e.g., adaptation to water stress) to use for 5750 planting (table 8.2) and consider using genetically improved seedling stock, particularly where 5751 species-specific insects or pathogens are a concern (e.g., white pine blister rust [Cronartium 5752 ribicola {J.C.Fisch}]). Overall, increasing diversity, including genotypic, phenotypic, species, 5753 and structural diversity, will help increase resilience to climate change and associated 5754 disturbances (Dymond et al. 2014, Halofsky et al. 2018). 5755 The distribution and abundance of some nonnative plant species will likely expand with 5756 climate change, because ecosystem disturbance and shifts in native species ranges will provide 5757 opportunities for their establishment (Halofsky and Peterson 2016). Some nonnative species are 5758 invasive, with characteristics that facilitate their expansion and dominance in a warmer climate, 5759 such as broad temperature tolerances and high dispersal ability (Hellman et al. 2008). Proactive 5760 management tactics such as early detection-rapid response for new invasions, incorporation of 5761 invasive species prevention in projects, and conducting outreach to educate employees and the 5762 public about invasives are often suggested by managers (Halofsky et al. 2016). Increasing 5763 collaboration among land owners and managers will facilitate effective control of invasives 5764 (Hellmann et al. 2008). Monitoring post-fire conditions will also help identify the presence of 5765 invasive species and increase chances to prevent establishment. 5766 5767 5768 Adapting Special Habitats Management to Climate Change in Southwest 5769 Washington 5770 5771 Riparian areas, wetlands, and groundwater-dependent ecosystems (GDEs) are important for 5772 many wildlife and botanical species in Southwest Washington, and increasing temperatures and 5773 altered hydrology will affect these habitat types (chapter 5). To retain species and maintain 5774 ecosystemDRAFT function, managers can work to restore ecosystems and reduce existing stressors 134

Hudec et al. Chapter 8

5775 (table 8.3). For example, maintaining or restoring stream channel form helps to increase 5776 hydrologic function and store water, which is beneficial for riparian and wetland vegetation, 5777 water quality, and aquatic habitat. To mitigate road impacts, road connectivity can be reduced, 5778 some roads decommissioned, and drainage redesigned to increase interception of precipitation 5779 and local retention of water (Peterson and Halofsky 2018) (table 8.3). In cooperation with the 5780 Washington Department of Fish and Wildlife, restoring and promoting American beaver 5781 populations could help to retain water and minimize extreme flows (Pollock et al. 2014, 2015). 5782 Monitoring will help managers understand changes underway in these habitats; adaptation 5783 workshop participants identified high priority locations for monitoring on the Gifford Pinchot 5784 National Forest, including South Prairie Fen, Lone Butte, and Skookum, Cayuse, Muddy, and 5785 Midway Meadows (table 8.3). Other adaptation strategies for these habitats include maintaining 5786 appropriate densities of native species, propagating drought-tolerant native species, and 5787 controlling nonnative species to increase resilience in a warmer climate (Peterson and Halofsky 5788 2018). 5789 Higher temperatures are likely to increase water stress for some plant species in subalpine 5790 and alpine plant communities (Hu et al. 2010). Increased monitoring will help to understand the 5791 ongoing effects of climate change on these plant communities (Peterson and Halofsky 2018), as 5792 well as the wildlife species that utilize them. Designating a long-term study area in alpine and 5793 subalpine habitats may be particularly effective for monitoring long-term changes and 5794 understanding human impacts (table 8.3). In particular, monitoring may help to identify climate 5795 change refugia, where current subalpine and alpine species may occur in the future and 5796 restoration may be most effective. Collecting seeds and creating a seedbank for high-elevation 5797 species (e.g., whitebark pine [Pinus albicaulis Engelm.]) will also help to ensure that seeds of 5798 native and local species are available for restoration efforts (table 8.3). 5799 As discussed in the previous section, increased area burned will reduce the extent of late- 5800 successional forest (chapter 4). Assessing where late-successional forests are most at risk to fire 5801 and insects will help prioritize actions such as fuel treatments and construction of fuel breaks. 5802 Connectivity of late-successional forest, which allows species to move in response to a changing 5803 climate (Mawdsley et al. 2009), may be of particular importance in Southwest Washington. 5804 5805 5806 Adapting Recreation and Ecosystem Services Management to Climate Change 5807 in Southwest Washington 5808 5809 Adapting recreation management to climate change in Southwest Washington will help ensure 5810 that recreation opportunities exist in the future. Increasing temperatures will have affect warm- 5811 weather recreation by extending the length of shoulder seasons (spring, autumn), and will affect 5812 snow-based recreation by reducing snowpack (chapter 6). Conflicts between recreation and 5813 wildlife may increase (e.g., snowmobile activity on Mt. Adams can affect wolverine [Gulo gulo 5814 L.] movement and population establishment). A general adaptation strategy to address these 5815 sensitivities is to transition recreation management to address changing recreation use patterns 5816 (Hand and Lawson 2017). 5817 A first step will be to conduct assessments to understand the changing patterns of use, 5818 estimating which areas and sites have increasing pressure in the shoulder seasons (table 8.4). 5819 Adjustments can then be made to increase the capacity of recreation sites that are showing 5820 increased DRAFTuse (e.g., campgrounds can be enlarged, and more fences, signs, and gates can be 135

Hudec et al. Chapter 8

5821 installed where necessary to direct use) (Hand and Lawson 2017). All-terrain vehicle use may 5822 increase during the shoulder seasons when trails are wet, which may damage trails. Managers 5823 may need to consider closing trails to some uses during the shoulder seasons to prevent damage. 5824 Roads will need to be managed different to accommodate year-round access, and timing of 5825 seasonal employees may also need to be adjusted. Managers can consider expanding wilderness 5826 areas (such as Trapper Creek) and other areas to accommodate increase in visitations by 5827 recreationists seeking solitude. Whenever possible, recreation managers can identify economic 5828 opportunities associated with increased recreation access and use to benefit communities (table 5829 8.4). 5830 Warming temperatures and changing disturbance patterns may alter the availability and 5831 timing of special forest products, leading to increased conflicts in uses (e.g., subsistence, 5832 cultural, traditional, and commercial uses) (chapter 7). Thus, managers may need to monitor and 5833 adaptively manage special forest products and the vegetation types in which they are found. 5834 Monitoring can help track changes over time to inform permitting and ensure sustainable harvest 5835 (table 8.4). 5836 Water-based recreation will likely become more popular as the public seeks relief from 5837 high summer temperatures (Hand and Lawson 2017). Thus, recreation impacts to riparian areas 5838 and lake shores may increase. Dispersed camp areas along waterways may need to be limited or 5839 managed more intensively to reduce impacts (table 8.4). Recreation managers can try to direct 5840 use to less sensitive areas. 5841 5842 5843 Conclusions 5844 5845 The SWAP vulnerability assessment and workshop resulted in a list of high-priority climate 5846 change adaptation strategies and tactics for natural resource management in the area. Many of 5847 the strategies and tactics were focused on increasing ecosystem resilience, although some were 5848 aimed at facilitating transition of ecosystems or management practices to a changing climate 5849 (e.g., transition recreation management to account for changing use patterns with climate 5850 change). Adaptation strategies and tactics that have benefits to more than one resource will 5851 generally have the greatest overall benefit (Peterson et al. 2011b). Management activities focused 5852 on reducing fuels and restoring hydrologic function are already standard practices on some state 5853 and federal lands in Southwest Washington, suggesting that many current resource management 5854 actions will also be appropriate in a changing climate. However, the locations where actions are 5855 implemented may be different or strategically targeted in the context of climate change. For 5856 example, fuel treatments in dry forest types may be targeted near high-value late-successional 5857 habitat, whose extent may decrease if fire frequency increases in the future. 5858 Implementation is the next challenging step for the SWAP (see chapter 9). Although 5859 implementing all adaptation options described here may not be feasible, managers can choose 5860 from the menu of strategies and tactics, and expand upon it in the future. Thus, these adaptation 5861 strategies and tactics can provide the basis for climate-informed management in the area. 5862 5863 5864 Literature Cited 5865 DRAFT 136

Hudec et al. Chapter 8

5866 Chmura, D.J.; Anderson P.D.; Howe G.T. [et al.]. 2011. Forest responses to climate change in 5867 the northwestern United States: ecophysiological foundations for adaptive management. 5868 Forest Ecology and Management. 261: 1121–1142. 5869 Dymond, C.C.; Tedder, S.; Spittlehouse, D.L. [et al.]. 2014. Diversifying managed forests to 5870 increase resilience. Canadian Journal of Forest Research. 44: 1196–1205. 5871 Goode, J.R.; Luce, C.H.; Buffington, J.M. 2012. Enhanced sediment delivery in a changing 5872 climate in semi-arid mountain basins: implications for water resource management and 5873 aquatic habitat in the northern Rocky Mountains. Geomorphology. 139: 1–15. 5874 Halofsky, J.E.; Peterson, D.L. 2016. Climate change vulnerabilities and adaptation options for 5875 forest vegetation management in the northwestern USA. Atmosphere. 7: 46. 5876 Halofsky, J.E.; Peterson, D.L.; Metlen, K.L. [et al.]. 2016. Developing and implementing 5877 climate change adaptation options in forest ecosystems: a case study in Southwestern 5878 Oregon, USA. Forests. 7: 268. 5879 Halofsky, J.E.; Peterson, D.L.; Prendeville, H.R. 2018. Assessing vulnerabilities and adapting 5880 to climate change in northwestern US forests. Climatic Change. 146: 89–102. 5881 Hamlet, A.F.; Lettenmaier, D.P. 2007. Effects of 20th century warming and climate variability 5882 on flood risk in the Western U.S. Water Resources. 43: W06427. 5883 Hand, M.S.; Lawson, M. 2017. Effects of climate change on recreation in the Northern 5884 Rockies. In: Halofsky, J.E.; Peterson, D.L., eds. Climate change and Rocky Mountain 5885 Ecosystems. Cham, Switzerland: Springer International Publishing: 169–188. 5886 Hellmann, J.J.; Byers, J.E.; Bierwagen, B.G.; Dukes, J.S. 2008. Five potential consequences 5887 of climate change for invasive species. Conservation Biology. 22: 534–543. 5888 Hu, J.; Moore, D.J.P.; Burns, S.P.; Monson, R.K. 2010. Longer growing seasons lead to less 5889 carbon sequestration by a subalpine forest. Global Change Biology. 16: 771–783. 5890 Independent Scientific Advisory Board [ISAB]. 2007. Climate change impacts on Columbia 5891 River Basin fish and wildlife. Rep. ISAB 2007-2. Portland, OR: Northwest Power Planning 5892 Council, Columbia River Basin Indian Tribes, and National Marine Fisheries Service. 5893 http://www.nwcouncil.org. (17 February 2017). 5894 Isaak, D.J.; Wollrab, S.; Horan, D.; Chandler, G. 2012. Climate change effects on stream and 5895 river temperatures across the northwest US from 1980–2009 and implications for salmonid 5896 fishes. Climatic Change. 113: 499–524. 5897 Littell J.S.; Oneil E.E.; McKenzie D. [et al.]. 2010. Forest ecosystems, disturbance, and 5898 climatic change in Washington State, USA. Climatic Change. 102: 129–158. 5899 Luce, C.; Morgan, P.; Dwire, K. [et al.]. 2012. Climate change, forests, fire, water, and fish: 5900 Building resilient landscapes, streams, and managers. Gen. Tech. Rep. RMRS-GTR-290. 5901 Fort Collins, CO: US Department of Agriculture, Forest Service, Rocky Mountain Research 5902 Station. 5903 Mantua, N.J.; Metzger, R.; Crain, P. [et al.]. 2011. Climate change, fish, and fish habitat 5904 management at Olympic National Forest and Olympic National Park. In: Halofsky, J.E.; 5905 Peterson, D.L.; O’Halloran, K.A.; Hawkins Hoffman, C., eds. Adapting to climate change at 5906 Olympic National Forest and Olympic National Park. Gen. Tech. Rep. PNW-GTR-844. 5907 Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research 5908 Station: 43–60. 5909 Mantua, N.J.; Raymond, C.L. 2014. Climate change, fish, and aquatic habitat in the North 5910 Cascade Range. In: Raymond, C.L.; Peterson, D.L.; Rochefort, R.M., eds. Climate change 5911 vulnerabilityDRAFT and adaptation in the North Cascades region, Washington. Gen. Tech. Rep. 137

Hudec et al. Chapter 8

5912 PNW-GTR-892. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific 5913 Northwest Research Station: 235–270. 5914 Mantua, N.J.; Tohver, I.; Hamlet, A. 2010. Climate change impacts on streamflow extremes 5915 and summertime stream temperature and their possible consequences for freshwater salmon 5916 habitat in Washington State. Climatic Change. 102: 187–223. 5917 Mawdsley, J.R.; O’Malley, R.; Ojima, D.S. 2009. A review of climate-change adaptation 5918 strategies for wildlife management and biodiversity conservation. Conservation Biology. 23: 5919 1080–1089. 5920 McCarthy, J.J.; Canziani, O.F.; Leary, N.A. [et al.], eds. 2001. Climate change 2001: 5921 impacts, adaptation, and vulnerability. Contribution of Working Group II to the Third 5922 Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, United 5923 Kingdom: Cambridge University Press. 5924 Millar, C.I.; Stephenson, N.L. 2015. Temperate forest health in an era of emerging 5925 megadisturbance. Science. 349: 823–826. 5926 Millar, C.I.; Stephenson, N.L.; Stephens, S.L. 2007. Climate change and forests of the future: 5927 managing in the face of uncertainty. Ecological Applications. 17: 2145–2151. 5928 Mote, P.W.; Parson, E.A.; Hamlet, A.F. [et al.] 2003. Preparing for climatic change: the water, 5929 salmon, and forests of the Pacific Northwest. Climatic Change. 61: 45–88. 5930 Muir, M.; Luce, C.H.; Gurrieri, J. [et al.]. 2018. Effects of climate change on hydrology, soil, 5931 and water resources. In: Halofsky, J.E.; Peterson, D.L.; Ho, J.J. [et al.], eds. Climate change 5932 vulnerability and adaptation in the Intermountain Region. Gen. Tech. Rep. RMRS-GTR-xxx. 5933 Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 5934 Station. xxx p. In press. 5935 Peterson, D.L.; Halofsky, J.E. 2018. Adapting to the effects of climate change on natural 5936 resources in the Blue Mountains, USA. Climate Services. 5937 https://doi.org/10.1016/j.cliser.2017.06.005. 5938 Peterson, D.L.; Halofsky, J.E.; Johnson, M.C. 2011a. Managing and adapting to changing fire 5939 regimes in a warmer climate. In: McKenzie, D.; Miller, C.; Falk, D., eds. The landscape 5940 ecology of fire. New York: Springer: 249–267. 5941 Peterson, D.L.; Millar, C.I.; Joyce, L.A. [et al.]. 2011b. Responding to climate change on 5942 national forests: a guidebook for developing adaptation options. Gen. Tech. Rep. PNW-GTR- 5943 855. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 5944 Research Station. 109 p. 5945 Pollock, M.M.; Beechie, T.J.; Wheaton, J.M. [et al.]. 2014. Using beaver dams to restore 5946 incised stream ecosystems. Bioscience. 64: 279–290. 5947 Pollock, M.M.; Lewallen, G.; Woodruff, K. [et al.], eds. 2015. The beaver restoration 5948 guidebook: working with beaver to restore streams, wetlands, and floodplains. Version 1.02. 5949 Portland, OR: U.S. Fish and Wildlife Service. 189 p. 5950 https://www.fws.gov/oregonfwo/Documents/BeaverRestGBv.1.02.pdf. (August 2, 2016). 5951 Rieman, B.E.; Isaak, D.J. 2010. Climate change, aquatic ecosystems and fishes in the Rocky 5952 Mountain West: implications and alternatives for management. Gen. Tech. Rep. GTR- 5953 RMRS-250. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky 5954 Mountain Research Station. 5955 Sohn, J.A.; Saha, S.; Bauhus, J. 2016. Potential of forest thinning to mitigate drought stress: a 5956 meta-analysis.DRAFT Forest Ecology and Management. 380: 261–273. 138

Hudec et al. Chapter 8

5957 Swanston, C.; Janowiak, M., eds. 2012. Forest adaptation resources: climate change tools and 5958 approaches for land managers. Gen. Tech. Rep. NRS-87. Newtown Square, PA: U.S. 5959 Department of Agriculture, Forest Service, Northern Research Station. 121 p. 5960 Young, M.K.; Isaak, D.J.; Spaulding, S. [et al.]. 2017. Effects of climate change on cold-water 5961 fish in the Northern Rockies. In: Halofsky, J.E.; Peterson, D.L., eds. Climate change and 5962 Rocky Mountain Ecosystems. Cham, Switzerland: Springer International Publishing: 37–58.

DRAFT 139

Hudec et al. Chapter 9

5963 Chapter 9: Conclusions 5964 5965 5966 Joanne J. Ho, David L. Peterson, and Jessica L. Hudec1 5967 5968 5969 The Southwest Washington Adaptation Partnership (SWAP) provided contributions to increase 5970 understanding of climate change vulnerabilities and assist climate change response in Southwest 5971 Washington and Gifford Pinchot National Forest. The effort synthesized the best available 5972 scientific information to assess climate change vulnerability for key resources of concern, 5973 develop recommendations for adaptation options, and catalyze a collaboration of land 5974 management agencies and stakeholders seeking to address climate change issues. Furthermore, 5975 the vulnerability assessment and corresponding adaptation options provided information to 5976 support Gifford Pinchot National Forest in implementing agency climate change objectives 5977 described in the National Roadmap for Responding to Climate Change (USDA FS 2010a) (see 5978 chapter 1). 5979 5980 5981 Relevance to U.S. Forest Service Climate Change Response Strategies 5982 5983 The SWAP process was directly relevant to the climate change strategy of the U.S. Forest 5984 Service (USFS). Information presented in this report is also relevant for other land management 5985 entities and stakeholders in the SWAP assessment area. This process can be replicated and 5986 implemented by any organization, and the adaptation options are applicable beyond USFS lands. 5987 Like previous adaptation efforts (e.g., Halofsky and Peterson 2017; Halofsky et al. 2011, 2018, 5988 in press; Raymond et al. 2014), a science-management partnership was critical to the success of 5989 the SWAP. Those interested in utilizing this approach are encouraged to pursue a partnership as 5990 the foundation for increasing climate change awareness, assessing vulnerability, and developing 5991 adaptation plans. 5992 5993 5994 Communication, Education, and Organizational Capacity 5995 5996 Organizational capacity to address climate change, as outlined in the USFS Climate Change 5997 Performance Scorecard (2011-2016) (USDA FS 2010b), required building institutional capacity 5998 in management units through information exchange and training for employees. Information 5999 sharing and training were built into the SWAP process through a one-day workshop in which 6000 resource managers and scientists presented results of the vulnerability assessment, including the 6001 effects of climate change on fish and aquatic habitat, vegetation, special habitats, recreation, and

1 Joanne J. Ho is a research economist with the University of Washington, School of Environmental and Forest Sciences, Seattle, WA; David L. Peterson is a senior research biological scientist, U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, 400 N 34th St., Suite 201, Seattle, WA 98103; Jessica L. Hudec is an ecologist, U.S. Department of Agriculture, Forest Service, Gifford Pinchot National Forest, Mt. Adams Ranger District, 2455DRAFT Hwy 141, Trout Lake, WA 98650. 140

Hudec et al. Chapter 9

6002 ecosystem services. The workshop introduced climate tools and techniques for assessing 6003 vulnerability and started the process of planning for adaptation. 6004 6005 6006 Partnerships and Engagement 6007 6008 Relationships developed through the SWAP process were as important as the products that were 6009 developed, because these relationships build the partnerships that are the cornerstone for 6010 successful agency responses to climate change. We built a partnership across federal, state, and 6011 county agencies, tribes, private industry, special interest groups, collaborative groups, and the 6012 University of Washington. This partnership will remain relevant for future forest planning efforts 6013 and restoration conducted by Gifford Pinchot National Forest in collaboration with other partners 6014 and stakeholders. By working with partners, the capability to respond effectively to climate 6015 change increases, especially through the use of an all-lands approach, which was an important 6016 context for the assessment. 6017 Climate change response is a relatively new and evolving aspect of land management, 6018 and the SWAP workshop provided an opportunity for participants to effectively communicate 6019 their professional experiences with respect to climate change and resource management in a 6020 collaborative and supportive environment. Because the workshop covered a broad range of 6021 topics, multidisciplinary large-group discussions resulted in conceptual breakthroughs across 6022 disciplines. 6023 6024 6025 Assessing Vulnerability and Adaptation 6026 6027 Elements 6 and 7 of the USFS Climate Change Performance Scorecard (2011-2016) required 6028 units to identify the most vulnerable resources, assess the expected effects of climate change on 6029 vulnerable resources, and identify management strategies to improve the adaptive capacity of 6030 national forest lands. The SWAP vulnerability assessment described the climate change 6031 sensitivity of multiple resources in Southwest Washington. Adaptation strategies and tactics 6032 developed for each resource area can be incorporated into resource-specific management plans. 6033 Science-management dialogue identified management practices that are useful for 6034 increasing resilience and reducing stressors and threats. Although implementing all adaptation 6035 options developed in the SWAP process may not be feasible, resource managers can still draw 6036 from the menu of options as needed. Some adaptation strategies and tactics can be implemented 6037 on the ground now, whereas others may require changes in management plans or policies, or 6038 become more appropriate as threats become more apparent. 6039 6040 6041 Science and Monitoring 6042 6043 Where applicable, SWAP products identified information gaps and uncertainties important to 6044 understanding climate change vulnerabilities and management influences on vulnerabilities. 6045 These information gaps could help determine where important monitoring and research would 6046 decrease uncertainties inherent to management decisions. In addition, current monitoring 6047 programs DRAFTthat provide information for detecting climate change effects and additional monitoring 141

Hudec et al. Chapter 9

6048 needs were identified for some resources in the vulnerability assessment. Working across 6049 multiple jurisdictions and boundaries will allow SWAP participants to potentially increase 6050 collaborative monitoring of climate change effects and effectiveness of adaptation actions. 6051 Scientific documentation in the assessment can also be incorporated into large landscape 6052 assessments such as the Gifford Pinchot National Forest land management plan, environmental 6053 analysis for National Environmental Policy Act (NEPA) projects, and specific project design 6054 criteria and mitigations. 6055 6056 6057 Implementation 6058 6059 Although challenging, implementation of adaptation options will gradually occur with time, 6060 often motivated by extreme weather and large disturbance events, and facilitated by changes in 6061 policies, programs, and land management plan revisions. It will be especially important for 6062 ongoing restoration programs to incorporate considerations for climate change adaptation to 6063 ensure effectiveness. A focus on thoroughly-vetted strategies may increase ecosystem function 6064 and resilience while minimizing implementation risk. Land management agencies, stakeholders, 6065 and Native American tribes working together will make implementation effective, particularly 6066 across boundaries. 6067 In many cases, similar adaptation options were identified for more than one resource 6068 sector, suggesting a need to integrate adaptation planning across multiple disciplines. Adaptation 6069 options that yield benefits to more than one resource are likely to have the greatest benefit 6070 (Halofsky et al. 2011, Peterson et al. 2011, Raymond et al. 2014). However, some adaptation 6071 options involve tradeoffs and uncertainties that need further exploration. Assembling an 6072 interdisciplinary team to tackle this issue will be critical for assessing risks and developing risk 6073 management options. Scenario planning may be a useful next step. 6074 Information in this assessment can be incorporated into everyday work through “climate- 6075 informed thinking,” assist in planning, and influence management priorities, such as public 6076 safety. Flooding, wildfires, and insect outbreaks may all be exacerbated by climate change, thus 6077 increasing hazards faced by federal employees and the public. Resource management can help 6078 minimize these hazards by restoring hydrologic function, reducing fuels, and modifying forest 6079 species composition. These management activities are commonplace, demonstrating that, in 6080 many cases, current resource management may already be climate informed. 6081 Implementation of adaptation actions may be limited by insufficient human resources, 6082 insufficient funding, and conflicting priorities. However, climate change-related effects are 6083 already apparent for some resource areas, such as changes in hydrologic regimes and area burned 6084 by wildfires. Thus, some adaptation options may be precluded and resources may be 6085 compromised if actions are not implemented soon. This creates an imperative for timely 6086 integration of climate change considerations as a component of resource management and 6087 agency operations. 6088 The climate change vulnerability assessment and adaptation approach developed by 6089 SWAP can be used by the USFS and other organizations in many ways. From the perspective of 6090 federal land management, this information can be integrated in the following aspects of agency 6091 operations: 6092 • Landscape management assessments/planning: The vulnerability assessment provides 6093 informationDRAFT on departure from desired conditions and best available science on climate 142

Hudec et al. Chapter 9

6094 change effects to resources. The adaptation strategies and tactics describe desired 6095 conditions and management objectives for inclusion in planning documents. 6096 • Resource management strategies: The vulnerability assessment and adaptation strategies 6097 and tactics can be used in forest resilience and restoration plans, conservation strategies, 6098 fire management plans, infrastructure planning, and State Wildlife Action Plans. 6099 • Project NEPA analysis: The vulnerability assessment provides best available science for 6100 documentation of resource conditions, climate change effects analysis, and alternatives 6101 development. Adaptation strategies and tactics provide mitigations and project design 6102 recommendations for specific locations. 6103 • Monitoring plans: The vulnerability assessment can help identify knowledge gaps that 6104 can be addressed by monitoring. 6105 • National forest land management plan revision process: The vulnerability assessment 6106 provides a foundation for understanding key resource vulnerabilities caused by climate 6107 change for the assessment phase of forest plan revision. Information from vulnerability 6108 assessments can be applied in assessments required under the USFS 2012 Planning Rule, 6109 describe potential climatic conditions and effects on key resources, and identify and 6110 prioritize resource vulnerabilities to climate change in the future. Climate change 6111 vulnerabilities and adaptation strategies can inform forest plan components such as 6112 desired conditions, objectives, standards, and guidelines. 6113 • Project design/implementation: The vulnerability assessment and adaptation strategies 6114 and tactics provide recommendations for mitigations and project design at specific 6115 locations. 6116 We are optimistic that climate change awareness, climate-informed management and 6117 planning, and implementation of climate change adaptation actions in the SWAP area will 6118 continue to evolve. We anticipate that within a few years: 6119 • Climate change will become an integral component of federal agency operations. 6120 • The effects of climate change on natural and human systems will be continually assessed. 6121 • Monitoring activities will include indicators to detect the effects of climate change on 6122 species and ecosystems. 6123 • Agency planning processes will provide more opportunities to manage across boundaries. 6124 • Restoration activities will be implemented in the context of the influence of a changing 6125 climate. 6126 • Management of carbon will be included in adaptation planning. 6127 • Organizational capacity to manage for climate change will increase within federal 6128 agencies and with local stakeholders. 6129 • Resource managers will implement climate-informed practices in long-term planning and 6130 management. 6131 This assessment provides the foundation for understanding potential climate change 6132 effects and implementing adaptation options that help reduce the negative impacts of climate 6133 change and transition resources to a warmer climate. We hope that by building on existing 6134 partnerships, the assessment will foster collaboration in climate change adaptation and resource 6135 management planning throughout Southwest Washington. 6136 6137 6138 Literature Cited 6139 DRAFT 143

Hudec et al. Chapter 9

6140 Halofsky, J.E.; Peterson, D.L., eds. 2017. Climate change vulnerability and adaptation in the 6141 Blue Mountains. Gen. Tech. Rep. PNW-GTR-939. Portland, OR: U.S. Department of 6142 Agriculture, Forest Service, Pacific Northwest Research Station. 6143 Halofsky, J.E.; Peterson, D.L.; Dante-Wood, S.K.; [et al.]. 2018. Climate change vulnerability 6144 and adaptation in the Northern Rocky Mountains. Gen. Tech. Rep. RMRS-GTR-374. Fort 6145 Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research 6146 Station. 6147 Halofsky, J.E.; Peterson, D.L.; Ho, J.J., eds. [in press]. Vulnerability of natural resources to 6148 climate change in South Central Oregon. PNW-GTR-xxx. Portland, OR: U.S. Department of 6149 Agriculture, Forest Service, Pacific Northwest Research Station. 6150 Halofsky, J.E.; Peterson, D.L.; O’Halloran, K.A.; Hoffman, C.H. 2011. Adapting to climate 6151 change at Olympic National Forest and Olympic National Park. Gen. Tech. Rep. PNW-GTR- 6152 844. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 6153 Research Station. 6154 Peterson, D.L.; Millar, C.I.; Joyce, L.A.; [et al.]. 2011. Responding to climate change in 6155 national forests: a guidebook for developing adaptation options. Gen. Tech. Rep. PNW-GTR- 6156 855. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 6157 Research Station. 6158 Raymond, C.L.; Peterson, D.L.; Rochefort, R.M., eds. 2014. Climate change vulnerability and 6159 adaptation in the North Cascades region. USDA Forest Service Gen. Tech. Rep. PNW-GTR- 6160 892. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest 6161 Research Station. 6162 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010a. National roadmap for 6163 responding to climate change. http://www.fs.fed.us/climatechange/pdf/roadmap.pdf. (9 6164 February 2017). 6165 U.S. Department of Agriculture, Forest Service [USDA FS]. 2010b. A performance scorecard 6166 for implementing the Forest Service climate change strategy. 6167 http://www.fs.fed.us/climatechange/pdf/performance_scorecard_final.pdf. (9 February 2017). 6168 U.S. Department of Agriculture, Forest Service [USDA FS]. 2012. 2012 Planning Rule. 6169 Washington, DC: U.S. Department of Agriculture, Forest Service. Federal Register. 77(68). 6170 115 p. http://www.fs.usda.gov/Internet/FSE_DOCUMENTS/stelprdb5362536.pdf. (9 6171 February 2017). 6172 DRAFT 144

1

Table 1.1—U.S. Forest Service policies related to climate change

Policy Description

Forest Service Strategic Framework Developed in 2008, the Strategic Framework is based on seven strategic for Responding to Climate Change goals in three broad categories: foundational, structural, and action. The seven goals are science, education, policy, alliances, adaptation, mitigation,

(USDA FS 2008) and sustainable operations.

Like the challenges themselves, the goals are interconnected; actions that achieve one goal tend to help meet other goals. The key is to coordinate approaches to each goal as complementary parts of a coherent response to climate change. All seven goals are ultimately designed to achieve the same end (the USFS mission): to ensure that Americans continue to benefit from ecosystem services from national forests and grasslands.

USDA 2010-2015 Strategic Plan In June 2010, the U.S. Department of Agriculture released the Strategic Plan that guides its agencies toward achieving several goals including Strategic (USDA FS 2010c) Goal 2—Ensure our national forests and private working lands are conserved, restored, and made more resilient to climate change, while enhancing our water resources. This goal has several objectives. Objective 2.2 is to lead efforts to mitigate and adapt to climate change. The performance measures under this objective seek to reduce greenhouse gas emissions by the U.S. agricultural sector, increase the amount of carbon sequestered on U.S. lands, and bring all national forests into compliance with a climate change adaptation and mitigation strategy. The USFS response to this goal includes the National Roadmap for Responding to Climate Change and Performance Scorecard.

National Roadmap for Responding Developed in 2011, the Roadmap integrates land management, outreach, to Climate Change and sustainable operations accounting. It focuses on three kinds of activities: assessing current risks, vulnerabilities, policies, and gaps in knowledge; (USDA FS 2010b) engaging partners in seeking solutions and learning from as well as

educating the public and employees on climate change issues; and managing for resilience in ecosystems and human communities through adaptation, mitigation, and sustainable consumption strategies.

Climate Change Performance To implement the Roadmap, starting in 2011, each national forest and Scorecard grassland began using a 10-point scorecard to report accomplishments and plans for improvement on 10 questions in four dimensions: organizational (USDA FS 2010a) capacity, engagement, adaptation, and mitigation. By 2015, each is expected

to answer “yes” to at least seven of the scorecard questions, with at least one “yes” in each dimension. The goal is to create a balanced approach to climate change that includes managing forests and grasslands to adapt to changing conditions, mitigating climate change, building partnerships across boundaries, and preparing employees to understand and apply emerging science.

2012 Planning Rule The 2012 Planning Rule is based on a planning framework that will facilitate adaptation to changing conditions and improvement in management based (USDA FS 2012) on new information and monitoring. There are specific requirements for addressing climate change in each phase of the planning framework, DRAFTincluding in the assessment and monitoring phases, and in developing, 2

revising, or amending plans. The 2012 Planning Rule emphasizes restoring the function, structure, composition, and connectivity of ecosystems and watersheds to adapt to the effects of a changing climate and other ecosystem drivers and stressors, such as wildfire and insect outbreaks. A baseline assessment of carbon stocks required in assessment and monitoring will check for measureable changes in the plan area related to climate change and other stressors.

Requirements of the Roadmap and Scorecard and requirements of the 2012 Planning Rule are mutually supportive and provide a framework for responding to changing conditions over time.

DRAFT 3

Table 1.2—Climate change vulnerability assessments on Forest Service lands (listed in chronological order)

Publication Project States Federal landsa Area Vulnerability Adaptation assessment? options? Million acres

Halofsky et Olympic Northwest Olympic NF, 1.6 Yes Yes al. 2011 Adaptation Washington Olympic NP Partnership

Rice et al. Assessment for Wyoming Shoshone NF 2.4 Yes No 2012 Shoshone National Forest

Littell et al. Assessment for California Tahoe NF, Inyo 1.2 Yes Yes 2012, Morelli eastern NF, Devils et al. 2012 California Postpile NM

Furniss et al. Watershed Nationwide 11 national forests NA Yes No 2013 Vulnerability across U.S. Assessments

Raymond et North Cascadia Northern Mt. Baker- 6.0 Yes Yes al. 2014 Adaptation Washington Snoqualmie NF, Partnership Okanogan- Wenatchee NF, North Cascades NP, Mt. Rainier NP

Janowiak et Northwoods Northern Chequamegon- 16.0 Yes No al. 2014 Climate Wisconsin, Nicolet NF, Change Michigan Ottawa NF Response Framework Project

Swanston et Climate Northern Chequamegon- 18.5 Yes No al. 2011, Change Wisconsin Nicolet NF 2016 Response Framework

Halofsky and Blue Mountains Oregon Malheur NF, 5.3 Yes Yes Peterson Adaptation Umatilla NF, 2017 Partnership Wallowa-Whitman NF

Hayward et Assessment for Alaska Chugach NF, 22.7 Yes No al. 2017 southcentral Kenai Peninsula Alaska

Halofsky et South Central Oregon Deschutes NF, 5.0 Yes Yes al. [in press Oregon Fremont-Winema b] Adaptation NF, Ochoco NF, Partnership Crooked River National Grassland, Crater Lake NP DRAFT 4

Halofsky et Northern Montana, 15 national 183 Yes Yes al. [in press Rockies northern forests, 3 national a] Adaptation Idaho, South parks Partnership Dakota, NW Wyoming

Halofsky et Intermountain Utah, 12 national 34.0 Yes Yes al. [in press Adaptation Nevada, forests, 22 c] Partnership southern National Park Idaho, NW Service units Wyoming

a NF=National Forest, NP=National Park, NM=National Monument.

DRAFT 5

Table 3.1—Summary of fish species of concern and climate vulnerability in the SWAP assessment area

Climate Species or run Range extent Population status/trenda vulnerability Comment Coho salmon Alaska through Depressed/stable Moderate ESA listeda California Chinook salmon Spring run Alaska through Depressed/stable Moderate ESA listed California Fall run Alaska through Depressed/stable to increasing Low ESA listed California Steelhead Summer run Alaska through Depressed/stable High ESA listed California Winter run Alaska through Depressed/stable Moderate ESA listed California Bull trout British Columbia Depressed/stable Moderate ESA listed through Oregon a Status and trend information are based on Ford et al. (2011), NOAA (2016), and NWFSC (2015).

DRAFT 6

Table 3.2—Summary of August mean stream temperatures in the SWAP assessment area during the baseline period and two future periods associated with the A1B emission scenario

< 8 oC 8–11 oC 11–14 oC 14–17 oC 17–20 oC > 20 oC ------Stream kilometers ------All lands 1980s 422 1,554 1,749 624 124 20 2040s 213 1,011 1,836 1,110 276 46 2080s 147 675 1,665 1,486 402 115

Forest Service lands 1980s 361 1,246 775 8 - - 2040s 174 875 1,136 203 1 - 2080s 134 583 1,203 463 3 -

DRAFT 7

Table 3.3—Summary of streamflow statistics relevant to fish populations in the SWAP assessment area, based on changes associated with the A1B emission scenario

All lands Forest Service lands Flow metrica Climate Number of Days increase Number of Days increase period daysb days Winter 95% flow 1980s 11.0 - 9.3 - 2040s 13.0 2.0 12.3 3.0 2080s 14.0 3.0 13.8 4.5

Percent Percent change change Peak flow 1980s - - - - 2040s - 9.6 - 13.0 2080s - 17.3 - 22.8

Cubic meters Percent Cubic Percent per second change meters per change second Mean summer flowc 1980s 2.7 - 1.7 - 2040s 1.6 -39.9 0.9 -44.6 2080s 1.1 -57.8 0.6 -64.9

Mean annual flow 1980s 5.3 - 2.6 - 2040s 5.5 3.3 2.7 3.1 2080s 5.5 3.9 2.7 3.6 a Stream reaches in network with mean summer flows greater than 0.006 m3s-1. b Refers to day of water year starting October 1. c Average flow across all reaches in the network.

DRAFT 8

Table 3.4—Summary of streamflow and temperature characteristics for the 584 kilometers of coho salmon habitat shown in fig. 3.7, based on changes associated with the A1B emission scenario. Values are stream kilometers; those in parentheses are percentages of the total

Number of high flow days

Stream metric Period <5 5-10 >10 Winter 95% flow 1980s 80 (14%) 162 (28%) 343 (59%) 2040s 37 (6%) 27 (5%) 521 (89%) 2080s - 40 (7%) 545 (93%) m3s-1 <0.034 0.034-0.085 >0.085 Summer flow 1980s 2.4 (1%) 41 (6%) 541 (93%) 2040s 4.6 (1%) 53 (9%) 526 (90%) 2080s 6.6 (1%) 55 (9%) 523 (89%) Stream kilometers <8 8-11 11-14 14-17 17-20 >20 August temperature 1980s 6.1 (1%) 73 (13%) 351 (60%) 107 (18%) 6 (1%) 40 (7%) 2040s - 30 (5%) 215 (37%) 284 (49%) 11 (2%) 45 (8%) 2080s - 18 (3%) 118 (20%) 375 (64%) 26 (5%) 46 (8%)

DRAFT 9

Table 3.5—Streamflow and temperature characteristics for the 359 kilometers of Chinook salmon habitat shown in figure 3.8, based on changes associated with the A1B emission scenario. Values are stream kilometers; those in parentheses are percentages of the total

Number of high flow days

Stream metric Period <5 5-10 >10 Winter 95% flow 1980s 65 (18%) 139 (39%) 155 (43%) 2040s 26 (7%) 23 (6%) 310 (86%) 2080s - 26 (7%) 333 (93%) m3s-1 <0.034 0.034-0.085 >0.085 Summer flow 1980s - 10 (3%) 349 (97%) 2040s - 20 (6%) 339 (94%) 2080s - 21 (6%) 338 (94%) Stream kilometers <8 8-11 11-14 14-17 17-20 >20 August temperature 1980s 4.1 (1%) 45 (13%) 241 (67%) 39 (11%) 6 (2%) 24 (7%) 2040s - 20 (6%) 133 (37%) 167 (46%) 12 (3%) 28 (8%) 2080s - 9 (2%) 83 (23%) 223 (62%) 15 (4%) 29 (8%)

DRAFT 10

Table 3.6—Summary of stream flow and temperature characteristics for the 901 kilometers of steelhead trout habitat shown in figure 3.9, based on changes associated with the A1B emission scenario. Values are stream kilometers, those in parentheses are percentages of the total

Number of high flow days

Stream metric <5 5-10 >10 Period Winter 95% flow 1980s 106 (12%) 160 (18%) 635 (70%) 2040s 37 (4%) 52 (6%) 812 (90%) 2080s - 40 (4%) 861 (96%) m3s-1 <0.034 0.034-0.085 >0.085 Summer flow 1980s 9 (1%) 94 (10%) 799 (89%) 2040s 30 (3%) 119 (13%) 752 (83%) 2080s 33 (4%) 127 (14%) 741 (82%) Stream kilometers <8 8-11 11-14 14-17 17-20 >20 August temperature 1980s 14 (2%) 167 (18%) 564 (63%) 108 (12%) 8 (1%) 40 (4%) 2040s 4 (1%) 58 (6%) 423 (47%) 359 (40%) 12 (1%) 45 (4%) 2080s 3 (0%) 29 (3%) 269 (30%) 520 (58%) 34 (4%) 46 (5%)

DRAFT 11

Table 3.7—Summary of stream flow and temperature characteristics for the 20 kilometers of bull trout spawning and rearing habitat in Gifford Pinchot National Forest, as shown in figures 3.10 and 3.11, based on changes associated with the A1B emission scenario. Values are stream kilometers, those in parentheses are percentages of the total

Number of high flow days

Stream metric Period <5 5-10 >10 Winter 95% flow 1980s - 2 (10%) 18 (90%) 2040s - - 20 (100%) 2080s - - 20 (100%) m3/s <0.034 0.034-0.085 >0.085 Summer flow 1980s - 4 (20%) 16 (80%) 2040s - 4 (20%) 16 (80%) 2080s - 4 (20%) 16 (80%) Stream kilometers <8 8-11 11-14 14-17 17-20 >20 August temperature 1980s - 20 (100%) - - - - 2040s - 8 (40%) 12 (60%) - - - 2080s - 6 (30%) 14 (70%) - - -

DRAFT 12

Table 4.1—Climate change vulnerability estimates for western Washington tree species, based on risk factor scores and overall scores, where higher scores indicate higher vulnerability (adapted from Devine et al. 2012)

Adaptive Insects Reproductive Habitat genetic and Overall Species Distribution capacity affinity variation diseases scorea

Pacific silver fir 19b 100 100 100 86 81

Subalpine fir 38 67 65 84 100 71

Engelmann spruce 100 67 54 84 25 66

Noble fir 59 67 50 100 31 61

Grand fir 57 67 4 50 92 54

Mountain hemlock 38 33 88 67 31 51

Alaska yellow-cedar 63 67 58 67 0 51

Western white pine 83 33 15 0 58 38

Douglas-fir 0 67 8 50 28 31

Bigleaf maple 35 0 15 50 47 29

Black cottonwood 63 0 23 34 20 28

Sitka spruce 57 33 39 0 3 26

Western redcedar 44 67 0 17 3 26

Western hemlock 13 0 39 34 25 22

Red alder 19 0 19 50 14 20

a Calculated by averaging the scores from the five risk factors, each ranging from 0-100.

b Numbers are relative within columns; magnitude has no inherent biological meaning.

DRAFT 13

Table 5.1—Special habitats and species of concern in Gifford Pinchot National Forest

Special habitats Animals Plants Forested ecosystems Late-successional Northern spotted owl Pseudocyphellaria rainierensis (cyanolichen) forest (Strix occidentalis) Nephroma occultum (cyanolichen) Marbled murrelet (Brachyramphus Dendriscocaulon intricatulum (cyanolichen) marmoratus) Lobaria linita (cyanolichen) Flying squirrel Leptogium cyanescens (cyanolichen) (Glaucomys sabrinus) Cetrelia cetrarioides (green algal lichen) Tholurna dissimilis (lichen) Lemondrop whiskers (Chaenotheca subroscida) Methuselah’s beard (Dolichousnea longissima) Tetraphis moss (Tetraphis geniculata)

Early-seral forest Elk Huckleberry (Cervus elaphus) (Vaccinium spp.) Black bear (Ursus americanus) Black-tailed deer (Odocoileus hemionus)

Bigleaf maple Puget Oregonian snail Bigleaf maple (Cryptomastix devia) (Acer macrophylla)

Subalpine parkland Clark's nutcracker Whitebark pine (Nucifraga columbiana) (Pinus albicaulis)

Forested ephemeral Bats Mosses and liverworts “pothole” wetlands Odonates Cottonwood (Populus trichocarpa) Sitka alder (Alnus viridis)

Aspen (upland) Aspen (Populus tremuloides)

Chinquapin Golden hairstreak butterfly Chinquapin (Habrodais grunus) (Chrysolepis chrysophylla)

Oregon white oak Western gray squirrel Oregon white oak (Sciurus griseus) (Quercus garryana)

Non-forested ecosystems Mardon skipper Pale blue-eyed grass Meadows (wet and dry) (Polites mardon) (Sisyrinchium sarmentosum) Elk Rosy owl clover (Cervus elaphus) (Orthocarpus bracteosus) Hairy-stemmed checker-mallow (Sidalcea hirtipes) Mountain buttercup (Ranunculus populago) Camas (Camassia quamash)

Rock outcrops Mountain goat Howell's fleabane (Oreamnos americanus) (Erigeron howellii) Barrett's beardtongue (Penstemon barrettiae) DRAFT 14

Alpine Wolverine Great Smoky Mountain sedge (Gulo gulo) (Carex proposita) Cascade red fox Western sweetvetch (Vulpes vulpes cascadensis) (Hedysarum occidentale) American pika Curved woodrush (Ochotona princeps) (Luzula arcuata) Lomatium spp.

Riparian, wetland, and groundwater-dependent ecosystems Riparian (including Neotropical birds Cold-water corydalis aspen/cottonwood) Amphibians (Corydalis aquae-gelidae) Harlequin duck Oregon bolandra (Histrionicus histrionicus) (Bolandra oregana) Barrow’s goldeneye Mountain moonwort (Bucephala islandica) (Botrychium montanum) Common goldeneye Dense sedge (Bucephala clangula) (Carex densa) Wood duck Large-awned sedge (Alix sponsa) (Carex macrochaeta) Oregon false goldenaster (Heterotheca oregana) Howell's rush (Juncus howellii) Luminous moss (Schistostega pennata) Aquatic lichens (Leptogium rivale, Peltigera hydrothyria) Wetlands, springs, and Oregon spotted frog Hairy-stemmed checker-mallow seeps (Rana pretiosa) (Sidalcea hirtipes) North American beaver (Castor Thin-leaf cottonsedge canadensis) (Eriophorum viridicarinatum) Northern silverpuffs (Microseris borealis) Flat leaf bladderwort (Utricularia intermedia)

Fens Amphibians Northern silverpuffs (Microseris borealis) Flat leaf bladderwort (Utricularia intermedia) Small bladderwort (Utriculatia minor) Small cranberry (Vaccinium oxycoccus)

DRAFT 15

Table 6.1—Summary of climate change assessment ratings for recreation by activity category

Activity Magnitude of Likelihood of Direct effects Indirect effects category climate effect climate effect

Warm-weather Moderate (+) High Warmer temperature (+) Increased incidence, area, and activities Higher likelihood of severity of wildfire (+/-) extreme temperatures (-) Increased smoke from wildfire (-)

Snow-based High (-) High Warmer temperature (-) winter activities Reduced precipitation as snow (-)

Wildlife Terrestrial Moderate Warmer temperature (+) Increased incidence, area, and activities wildlife: low (+) Higher incidence of low severity of wildfire (terrestrial Fishing: stream flow (fishing -) wildlife +/-) moderate (-) Reduced snowpack Reduced cold-water habitat, (hunting -) incursion of warm-water tolerant species (fishing -)

Gathering Low (+/-) Moderate Warmer temperature (+) More frequent wildfires (+/-) forest products Higher severity wildfires (-)

Water-based Moderate (+) Moderate Warming temperatures (+) Lower stream flows and reservoir activities, not Higher likelihood of levels (-) including extreme temperatures (-) Earlier season low-flows (-) fishing Increased incidence of water quality degradation (e.g., algal blooms) (-)

Note: Positive (+) and negative (-) signs indicate expected direction of effect on overall benefits derived from recreation activity.

DRAFT

Table 7.1—Special forest products in permits sold and revenue dollars (nominal) for Gifford Pinchot National Forest, fiscal years (FY) 2014 and 2015. The upper three rows of data are for contract agreements rather than individual permits Products FY 2015 FY 2014 ------$ ------Permits------$ ------Permits-- Boughs (contract) 293,813 17 459,379 12 Firewood (contract) 1,526 3 6,752 9 Transplants (contract) 930 1 ------

Beargrass 131,685 2,561 111,275 2,079 Salal 83,010 1,415 58,060 962 Mushrooms 87,316 1,159 64,433 790 Berries 31,890 656 73,155 1,198 Transplants 1,050 3 Cuttings 800 37 1,852 1 Cones ------Fiber ------Boughs ------40 2 Firewood 18,865 982 23,420 880 Poles/posts 220 10 436 18 Christmas trees 4,500 920 11,365 2,467 Christmas trees/vendor 7,978 1,773 Cuttings ------661 34 Transplants 850 4 40 2 Yew billets/staves ------Bark/stumps/conks ------Cones ------Moss ------Restitution ------Misc. non-convertible 1,900 113 2,426 126 Misc. convertible 294 2 1,472 2 Mushrooms 64,340 3,134 33,854 1,729 Edible ferns 1,920 92 2,560 128 Total 731,843 12,879 852,230 10,442 DRAFT 16

Table 8.1—Fish and aquatic habitat adaptation options for Southwest Washington Sensitivity to climate Adaptation strategy Adaptation tactic change Increased flood frequency Increase spawning habitat • Restore stream and floodplain complexity and higher peak flows may resilience by restoring stream • Increase protection of alternate spawning reduce egg-fry survival for and floodplain structure and habitat fall spawners and yearling processes • Consider removing or modifying natural parr winter survival. barriers to increase spawning habitat • Increase use of logjams where feasible • Increase bank and channel stability* Increase habitat resilience by • Designate and restore natural flood plain reducing threats from roads and boundaries infrastructure in the floodplain • Increase floodplain habitat • Remove infrastructure from floodplains • Disconnect roads from streams • Reduce road density near streams • Increase culvert capacity • Increase side channel habitat and large wood for parr winter survival* Lower low flows will reduce Increase aquatic habitat • Increase off-channel habitat and protect fish habitat quality resilience to low summer flows refugia in side channels and channels fed by wetlands • Protect wetland-fed streams that maintain higher summer flows • Design channels at stream crossings to provide a deep thalweg for fish passage during low-flow periods • Increase deep water habitat and channel morphology* • Reduce width to depth ratios to reduce solar radiation in stream* Manage upland vegetation to • Increase forest cover to retain snow and retain water and snow in order decrease snow melt to slow spring snowmelt and • Restore mid-and high-elevation wetlands runoff that have been altered by land use

Manage riparian vegetation to • Plant trees* optimize shade to streams* • Maintain or enhance shade to streams* • Increase sinuosity in channels* • Eliminate human disturbances affecting width to depth ratio*

Protect existing hyporheic • Avoid activities that disrupt flows (e.g., flows* roads)* • Identify locations of hyporheic flows* Lower low flows will Increase in-stream flows with • Increase efficiency of irrigation techniques increase pre-spawn dry-season water conservation • Reduce summer withdrawals on federal mortality for summer run to reduce withdrawals lands and stream-type salmon and • Consider alternative water supplies for steelhead federal lands to retain in-stream flows • Coordinate with downstream partners on water conservation education • Restore beaver habitat and colonies* • Investigate and quantify connectivity between groundwater and stream flows* DRAFT 17

Warmer stream Increase habitat resilience for • Increase habitat and refugia in side temperatures will reduce cold-water fish by restoring channels thermal heterogeneity in structure and function of • Protect wetland-fed streams that maintain streams and increase streams higher summer flows thermal stress on many life • Restore structure and heterogeneity of stages of fish stream channels • Reconnect floodplains and side channels to improve hyporheic and base flow conditions • Remove dikes and levees • Restore and protect riparian vegetation • Manage livestock grazing to restore ecological function of riparian vegetation and maintain streambank conditions • Reduce high road densities that are intercepting subsurface stream flows* Increase understanding of • Conduct field experiments of fish- thermal tolerance of fish temperature relationships for multiple species species and regions • Monitor changes in stream temperature and fish distributions • Re-evaluate and update water temperature standards (both values and indices) • Manage fishing to reduce stress to fish during critical times* • Evaluate nonnative species that might expand and plan ahead for management* • Tailor restoration actions to benefit native species* • Increase public education on the issue (brochure, flyer, web, signage)* Increase understanding of • Identify and inventory cold-water refugia, thermal heterogeneity in springs, and groundwater input to springs streams and cold-water refugia • Identify seasonal refugia (winter and summer) • Research the influences of lakes, reservoirs, and groundwater on stream temperatures • Research fish use of thermal refugia Warmer stream Increase resilience of native fish • Survey and map nonnative species temperatures may favor species by reducing barriers to • Combine nonnative mapping with nonnative fish species native species and removing information on migration barriers nonnative species • Consider information from surveys of warmer basins farther south as indicators of vulnerability • Remove or control nonnative fish species • Assess migration barriers and potential habitat for native species • Remove barriers to fish passage where this will not increase threats from nonnative species • Maintain or construct barriers to prevent spread of nonnative species Warmer stream Increase population resilience • Increase public education to eliminate temperatures may create by increasing fish health disease vectors more favorable conditions • Direct treatment or removal of infected fish for diseases and parasites • Survey fish health conditions • Collaborate and standardize health survey methods among agencies DRAFT• Consider changes in hatchery practices 18

Warmer summer stream Monitor changes in aquatic food • Assess food webs for baseline data temperatures may alter web dynamics • Monitor food web dynamics for changes aquatic food web dynamics with warming • Compile existing research for data gaps Increased flooding and Manage and reduce sediment • Evaluate road system for sediment input landslides will increase generated by roads • Reduce sediment input to streams by sedimentation in streams replacing (or resizing) culverts, and relocating and decommissioning roads and increase frequency of ditch release culverts Identify hillslope landslide • Use Forest Service regional landslide risk hazards and at risk roads* model to identify and prioritize areas most at risk* • Identify infrastructure (roads, trails, recreation sites) at risk, and prioritize roads and other infrastructure for removal, modification, or relocation* Sedimentation in streams Reduce sedimentation • Include climate change projections in will increase as fire area and associated with erosion, fire, identification of potential areas for stream fire severity increase and trails. bank and upland erosion • Inventory disturbed areas for riparian and upland vegetation restoration • Manage fire and fuels with thinning and prescribed fire to reduce fire severity and extent • Restore and revegetate burned areas to store sediment and maintain channel geomorphology • Develop a geospatial layer of debris flow potential for pre-fire planning Identify hillslope landslide • Link stream inventory with topographic, hazard areas and at-risk roads geomorphic, and vegetation layers to prior to wildfires and as part of assess existing hazard and risk fire planning * Indicate adaptation strategies and tactics added in the workshop to existing Climate Change Adaptation Library strategies and tactics (http://adaptationpartners.org/library.php). DRAFT 19

Table 8.2—Forest vegetation adaptation options for Southwest Washington Sensitivity to climate Adaptation strategy Adaptation tactic change Climate change may Increase genotypic and • Actively migrate genetic material (and sufficiently alter conditions phenotypic diversity mycorrhizal associates) through movement such that genetic material is of seedlings no longer optimal for a given • Collect seed from multiple seed zones to site. increase seedbank and analyze for suitability within zone of interest • Identify suitable seed within existing zones; collect seed from trees that have desirable traits Climate change will likely Plan for increased area burned • Incorporate climate change in management result in increased area plans burned by wildfire. • Actively manage fire in grand fir and Douglas-fir vegetation zones • Expect increased area in early seral conditions • Consider management of fuels (all facets) to facilitate transition to new fire regimes in dry forest types Increased wildfire area Actively manage, protect and • Continue to aggressively suppress fires in burned may decrease area develop late seral structure and western hemlock and Pacific silver fir of late seral forest diversity forests • Increase active management in early and mid-seral western hemlock and Pacific silver fir forest stands to facilitate development of late seral structure • Analyze potential for minimizing risk of stand-replacing fire in late seral forest

DRAFT 20

Table 8.3—Special habitats adaptation options for Southwest Washington Sensitivity to climate Adaptation strategy Adaptation tactic change Increased summer Maintain current native • Conduct monitoring to understand changes temperatures and drought vegetation in alpine and underway stress may cause changes subalpine habitats • Collect seeds and create a seedbank for in herbaceous vegetation in high elevation species alpine and subalpine • Designate a natural study area in habitats. alpine/subalpine habitat to more effectively monitor long-term changes and understand human impacts Increase connectivity of habitat • Partner with adjacent land owners and islands managers to promote connectivity • Increase connectivity around and across Highway 12 Tree encroachment will Manage alpine wilderness • Establish burn plan for wilderness areas likely reduce area of alpine areas meadows. Reduced snowpack, Mitigate changes to the • Mitigate road impacts; eliminate summer precipitation, and hydrologic regime to retain unnecessary roads and their impacts to changing groundwater species and ecosystem function wetlands recharge and discharge will in riparian areas, wetlands and • Redesign road drainage to increase water result in shifting plant groundwater-dependent retention; reduce runoff and increase species composition and ecosystems infiltration reduced habitat quality in • Work with the Washington Department of riparian areas, wetlands and Fish and Wildlife to restore beaver groundwater-dependent populations ecosystems. • Conduct monitoring to understand changes underway (e.g., in South Prairie Fen, Lone Butte, and Skookum, Cayuse, Muddy, and Midway Meadows; see chapter 5) • Create native seed bank for these ecosystems • Conduct stream restoration (e.g., connect floodplains and create side channels) Late seral forest may be lost Protect late-seral forest • Assess where late-seral forests are most at as a result of fire, drought risk to fire and insects stress, and insect outbreaks. • Reassess and revise Late Successional Reserve Assessments

DRAFT 21

Table 8.4—Recreation and ecosystem services adaptation options for Southwest Washington Sensitivity to climate Adaptation strategy Adaptation tactic change Extended shoulder seasons Anticipate increased and • Identify individual sites with increasing lead to overlap of seasonal shifting seasonal recreation pressure in shoulder seasons recreation uses traditionally patterns and increase resilience • Identify emerging recreation opportunities limited to summer or winter, to change and shift marketing to take advantage of as well as different, and opportunities to benefit communities potentially more, • Identify possible economic benefits of recreational opportunities. increasing shoulder season access • Manage roads for potentially year round access • Increase Forest Service and partner staff presence in areas where motorized uses are increasing; leverage partnerships to increase volunteer presence. • Allow overlapping, staggered tours of seasonal employees Shifts in availability and Manage product harvest timing, • Monitor and adaptively manage special timing of special forest location, and user-types forest products and related vegetation products may lead to types (e.g., salal, beargrass) conflicts in uses (e.g., • Track changes in use over time to inform subsistence, heritage, and permitting for sustainable harvest levels commercial uses). • Assess shifting use patterns for cross- resource impacts (e.g., wildlife) and direct use away from highly vulnerable areas • Determine effects from increased access with longer shoulder seasons and target staffing to high-demand areas Water demands from Manage recreation use and • Manage riparian areas to keep water cool recreation may impinge on infrastructure to minimize for fish water needs of other impacts associated with • Locate facilities and infrastructure based on resource areas, namely fish. changes in human use anticipated future recreation demands • Inventory and track the heaviest use and/or damage in dispersed camp areas; enforce occupancy limits • Prevent expansion of dispersed camp areas by placing rocks or blocking access

DRAFT 22

1

Figure 1.1—Project area for the Southwest Washington Adaptation Partnership. Note that the Yakama Reservation is not included in the assessment area.

DRAFT

2

Figure 2.1—Huckleberry field in Gifford Pinchot National Forest with Mt. Adams in the background. (Photo by Eric Erlenbusch)

Figure 2.2—Mt. St. Helens from the north, before (left) and after (right) the 1980 eruption, with Spirit Lake in the foreground. (Photos by U.S. Geological Survey)

DRAFT

3

Figure 2.3—Annual historical temperature for Gifford Pinchot National Forest. The top of the gray range represents mean annual maximum temperature (i.e., mean of monthly maximum temperatures), the bottom of the range represents mean annual minimum temperature (i.e., mean of monthly minimum temperatures), and the black line represents mean annual temperature. Data source: PRISM (Daly et al. 2001). Analysis by J. Miller, U.S. Forest Service, Pacific Northwest Region.

Figure 2.4—Annual historical precipitation for Gifford Pinchot National Forest. Data source: PRISM (Daly et al. 2001). Analysis by J. Miller, U.S. Forest Service, Pacific Northwest Region.

DRAFT

4

Figure 2.5—Climate assessment area in southwest Washington.

DRAFT

5

Figure 2.6—Projected future mean annual temperature for southwest Washington from three global climate models (HadGEM2-ES, CSIRO-Mk3.6.0, and NorESM1-M) under Representative Concentration Pathway 8.5 for three future time periods. Historical mean annual temperature (derived from PRISM data) is shown with a dashed black horizontal line.

Figure 2.7—Historical (PRISM) and projected future mean monthly temperature for southwest Washington from three global climate models (HadGEM2-ES, CSIRO-Mk3.6.0, and NorESM1- M) under Representative Concentration Pathway 8.5 for 2070-2099.

DRAFT

6

Figure 2.8—Projected future mean annual precipitation for southwest Washington from three global climate models (HadGEM2-ES, CSIRO-Mk.3.6.0, and NorESM1-M) under Representative Concentration Pathway 8.5 for three future time periods. Historical mean annual precipitation (derived from PRISM data) is shown with a dashed black horizontal line.

Figure 2.9—Historical (PRISM) and projected future mean monthly precipitation for southwest Washington from three global climate models (HadGEM2-ES, CSIRO-Mk3.6.0, and NorESM1- M) under Representative Concentration Pathway 8.5 for 2070-2099.

DRAFT

7

Figure 3.1—SWAP assessment area showing perennially flowing streams and areas where anadromous fish access is blocked by natural or human-caused factors.

DRAFT

8

Figure 3.2—NorWeST August mean stream temperature scenarios interpolated from 1,258 summers of monitoring data at 344 unique stream sites across the 4,500 km of streams in the assessment area. Map panels show conditions for baseline (a: 2000s) and late century scenarios (b: 2080s). High-resolution digital images of these maps and ArcGIS databases with reach-scale predictions are available at the NorWeST website (http://www.fs.fed.us/rm/boise/AWAE/projects/NorWeST.html).

Figure 3.3—Decadal river temperature trends for 1976-2015, estimated from long-term monitoring records available for July and August in the SWAP assessment area. A cooling trend at the Wind River site during August is associated local springs and flow regulation at the Carson National Fish Hatchery. Trend estimates are a subset of those reported for a regional river temperature trend analysis in Isaak et al. (2018).

DRAFT

9

Figure 3.4—Frequency of days when winter high flows are among the highest 5 percent of the year for the 1980s (panel a) and 2080s (panel b) based on the A1B emission scenario.

Figure 3.5—Percent change in peak flows projected by the Safeeq et al. (2015) model during the 2040s (panel a) and 2080s (panel b) relative to the 1980s baseline and the A1B emission scenario.

DRAFT

10

Figure 3.6—Percentage decline in mean summer flows predicted by the VIC model during the 2040s (panel a) and 2080s (panel b) relative to the 1980s baseline and the A1B emission scenario.

DRAFT

11

Figure 3.7—Distribution of coho salmon habitat assessed in the SWAP assessment area. Geospatial data are based on the most recent fish habitat database in Gifford Pinchot National Forest.

DRAFT

12

Figure 3.8—Distribution of Chinook salmon habitat in the SWAP assessment area. Geospatial data are based on the recent fish habitat database for Gifford Pinchot National Forest.

DRAFT

13

Figure 3.9—Distribution of steelhead trout habitat assessed in the SWAP assessment area. Geospatial data are based on the most recent fish habitat database in Gifford Pinchot National Forest.

DRAFT

14

Figure 3.10—Distribution of bull trout habitat assessed in the SWAP assessment area. Geospatial data are based on the most recent fish habitat database in Gifford Pinchot National Forest and on designations of critical habitat by the U.S. Fish and Wildlife Service.

DRAFT

15

Figure 3.11—Summer temperatures (ºC) in two streams that support spawning and rearing by bull trout in Gifford Pinchot National Forest. Stream reaches colored green may become too warm for spawning and rearing in the future; panel a shows 1980s, panel b shows 2080s, based on NorWeST and the A1B emission scenario.

DRAFT

16

Figure 4.1—Vegetation assessment area in Southwest Washington.

DRAFT

17

Figure 4.2—Vegetation profile through the Cascade Range and Columbia River Gorge (Troll 1955).

DRAFT

18

Figure 4.3—Average annual precipitation for Southwest Washington (1981-2010, PRISM data).

DRAFT

19

Figure 4.4—Potential vegetation zones (series) in Gifford Pinchot National Forest, organized by gradients of moisture and temperature (Topik 1989).

DRAFT

20

Figure 4.5—MC2 projections for a) wildfire, b) net primary productivity, and c) aboveground carbon. Historical conditions and three global climate models (CSIRO, HadGEM and NorESM) under the RCP 8.5 emission scenario are displayed.

DRAFT

21

Figure 4.6—MC2 projections of historical (1979-2008) and future vegetation composition for three global climate models (CSIRO, HadGEM, and NorESM) under the RCP 8.5 emission scenario.

DRAFT

22

Figure 4.7—Future vegetation composition for three global climate models (CSIRO, HadGEM, and NorESM) under the RCP 8.5 emission scenario, as projected by the MC2 model (top row) and climate-informed state-and-transition simulation models (STSMs), averaged across all Monte Carlo simulations (bottom row).

DRAFT

23

Figure 4.8—Areas of agreement in future potential vegetation type projected by MC2 for three global climate models (CSIRO, HadGEM, and NorESM) under the RCP 8.5 emission scenario. Colors represent areas of agreement. Box A shows 79 percent agreement in future vegetation type among the three models for 2010-2039. Box B shows 71 percent agreement in future vegetation type among the three models for 2040-2069. Box C shows 60 percent agreement in future vegetation type among the three models for 2070-2099.

DRAFT

24

Figure 4.9—Trends and variation in vegetation projections from climate-informed state-and- transition simulation models (cSTSMs). Note the difference in scale of the y-axes. Variation in trends shown is among cSTSM Monte Carlo simulations for three global climate models (CSIRO, HadGEM, and NorESM) under the RCP 8.5 emission scenario.

DRAFT

25

Figure 4.10—The alpine zone on Mount St. Helens supports a mixture of subshrubs, sedgeturf, and krummholz trees. (Photo by U.S. Forest Service)

Figure 4.11—Trampling by hikers poses a threat to alpine and parkland vegetation on the Mt. Adams South Climb route. (Photo by Paloma Ayala)

DRAFT

26

Figure 4.12—The Cold Springs Fire (2008) resulted in high mortality in mountain hemlock, subalpine fir, and grand fir zone forests on the south side of Mt. Adams. Much of the area burned again in the Cascade Creek Fire (2012) and Cougar Creek Fire (2015). (Photo by U.S. Forest Service)

DRAFT

27

Figure 4.13—Beargrass blooms beneath a canopy of Pacific silver fir trees. (Photo by U.S. Forest Service)

DRAFT

28

Figure 4.14—Forests of the grand fir zone in Gifford Pinchot National Forest suffer from high levels of mortality and heavy fuel loadings as a result of insects, disease, and fire exclusion. (Photo by U.S. Forest Service)

DRAFT

29

Figure 5.1—Late-Successional Reserves in Gifford Pinchot National Forest.

DRAFT

30

Figure 5.2—Projected trends from climate-informed state-and-transition simulation models (cSTSM) for large diameter, multi-storied forests and grass/forb and post disturbance forests in (a) high-elevation forests of the subalpine parkland, mountain hemlock, and silver fir zones and (b) low-elevation forests. Variation in trends shown is from an ensemble of cSTSM Monte Carlo simulations for three different climate scenarios: CSIRO, HadGEM, and NorESM. Note the difference in scale of y-axes. a)

b)

DRAFT

31

Figure 5.3—Alder Lake fire (2015) and associated fire management activities affected old growth and late-successional forest in the Mineral Late-Successional Reserve (LSR), Gifford Pinchot National Forest. Known nesting sites for marbled murrelet and northern spotted owl existed in Mineral LSR prior to the fire. Current status of nesting sites is unknown. (a) Fire smolders in dead and down coarse woody debris, thick organic soil horizons, and recently burned or felled large-diameter trees under a closed, multi-story canopy. Large amounts of fuel in late- successional forests can burn with high intensity and long duration, affecting various soil attributes and damaging tree roots. (Photo by Fred Way, Inciweb) (b) Professional tree fallers and heavy equipment construct containment lines. Old-growth and late-successional forests were affected by new fireline construction and improvements to roads, including the falling of large- diameter trees. (Photo by the National Interagency Fire Center) a)

b)

DRAFT

32

Figure 5.4—Conifer encroachment in huckleberry fields is a management concern in Gifford Pinchot National Forest. Two recent restoration projects in the Sawtooth and Polepatch traditional huckleberry picking areas have addressed conifer encroachment through mechanical treatments and prescribed fire. (a) Site conditions are monitored in the Sawtooth huckleberry fields prior to prescribed burning. (b) Firefighters observe fire behavior during a prescribed burn in the Sawtooth huckleberry fields. (Photos by U.S. Forest Service) a)

b)

DRAFT

33

Figure 5.5—Quaking aspen regeneration is abundant two years after the 2008 Cold Springs Fire. (Photo by U.S. Forest Service)

Figure 5.6—Subalpine parkland, alpine and subalpine meadows, rock outcrops, and alpine peaks in Goat Rocks Wilderness. Trees advance into meadows as conditions permit. (Photo by Dean Myerson)

DRAFT

34

Figure 5.7—Whitebark pine seedling on Mt. Adams germinated on bare, rocky soils. (Photo by U.S. Forest Service)

DRAFT

35

Figure 5.8—Meadows in Gifford Pinchot National Forest.

DRAFT

36

Figure 5.9—This wet subalpine meadow is part of the Bird Creek Meadows system on the southeast side of Mt. Adams. A serpentine creek is visible here in early August. Subalpine meadows like Bird Creek and Gotchen Meadows on Mt. Adams and Snowgrass Flats in Goat Rocks Wilderness are popular destinations for wildflower viewing. Altered hydrologic regimes in a warmer climate may alter species composition in meadows. Seasonal drying could increase foot traffic in the meadows and make plants more vulnerable to trampling. (Photo by U.S. Forest Service)

DRAFT

37

Figure 5.10—Mountain goat on a rock outcrop known as Sleeping Beauty, Gifford Pinchot National Forest, Mt. Adams Ranger District. (Photo by Dean Myerson)

Figure 5.11—Dwarfed and gnarled stature of krummolz at treeline on Mt. Adams provides evidence of harsh growing conditions in these high-elevation environments. (Photo by U.S. Forest Service)

DRAFT

38

Figure 6.1—Recreation sites (all types) in National Forest System land in Southwest Washington (recreation sites in Mt. Rainier National Park are not included here). Some sites may support more than one type of developed recreation.

DRAFT

39

Figure 6.2—Direct and indirect effects of climate on recreation decisions (from Hand and Lawson 2017).

DRAFT

40

Figure 6.3—More intense rain-on-snow events, such as the 2006-2007 floods in Southwest Washington, can cause significant damage to infrastructure and access. (Photo by U.S. Forest Service)

Figure 6.4—Locations with adequate snow for winter recreation (e.g., Sno-Parks) may face increases in use and a diminished recreational experience. (Photo by U.S. Forest Service)

DRAFT

41

Figure 6.5—In addition to sufficient snow cover, backcountry skiing requires scenic vegetation and solitude—an interaction of physical, biological, and social factors. (Photo by U.S. Forest Service)

DRAFT

42

Figure 7.1—Cut and sold timber volumes for Gifford Pinchot National Forest, 1920-2015.

Figure 7.2—Special forest products revenue (US$, nominal) for Gifford Pinchot National Forest, 2005-2015.

DRAFT

43

Figure 7.3—Baseline carbon stock for national forests within the Pacific Northwest Region from 2005 to 2013.

Figure 7.4—Carbon stored in harvested wood products (HWP) still in use and in solid waste disposal sites (SWDS) for the U.S. Forest Service Pacific Northwest Region (Butler et al. 2014). This carbon is not included in baseline carbon estimates for individual national forests (fig. 7.3), because it is typically located off site.

DRAFT

44

Figure 7.5—Percentage of smoke-impacted summer days at the IMPROVE station in White Pass, Washington, 2000-2013; vertical bars represent 1 standard deviation (data from Jim Miller).

DRAFT

1 Table for box 4.1—Net percent change in occurrence (1990-2010) of lichen species across 2 Southwest Washington and northwest Oregon, on west slope of the Cascade Range 3 (Gifford Pinchot, Mount Hood, and Willamette National Forests; n = 90 plots). The 4 number of times plots were sampled varies by location and species. 5 Species Change Percent Warm climate low-elevation species Evernia prunastri 64 Melanexelia subaurifera 0 Parmelia sulcata 13 Physcia adsendens 0 Ramalina farinacea 67

Cool climate montane species Platismatia glauca 3 Platismatia herreii 19 Platismatia stenophylla 22 Hypogymnia apinnata 5 Hypogymnia inactiva 42

Cold climate high-elevation species Alectoria sarmentosa -1 Cetraria merrellii -54 Cetraria platyphylla 39 Parmeliopsis ambigua -28 Parmeliopsis hyperoptera 19 6

DRAFT 1

7 Box 4.1—Lichen monitoring 8 9 Temperature, humidity, and precipitation influence the distribution of lichen species across the 10 landscape (Ellis 2007, Gauslaa 2014, Geiser and Neitlich 2006, Root et al. 2015). Lichen species 11 often occur in communities associated with specific climatic conditions, ranging from maritime 12 to high elevation to interior continental environments. Some lichen species are so sensitive to 13 temperature or precipitation that they live in a very narrow range of conditions (Ellis 2007, 14 Glavich et al. 2005, Martin 2005), with the potential for changes in abundance and distribution if 15 those conditions change. 16 The U.S. Forest Service Forest Inventory and Analysis (FIA) program and Pacific 17 Northwest Region Air Resources Management (ARM) program have monitored lichens since the 18 1990’s (Geiser and Neitlich 2006, USDA FS 2002). The ARM program, using lichen survey data 19 from the FIA permanent plot grid and ARM off-grid plots, has produced air quality and climate 20 models for Oregon and Washington (Root et al. 2015). The ARM program re-surveys this 21 network of field sites every 10 years, and some sites have now been surveyed over 20 years. 22 The 20-year lichen data from Gifford Pinchot, Mount Hood, and Willamette National 23 Forests were analyzed for correlations with climatic data across Southwest Washington and 24 Northwest Oregon. Indicator species analyses (ISA) (McCune and Grace 2002) of the 1990s 25 lichen data identified baseline climate-zone indicator species for warm low-elevation, cool 26 montane, and cold high-elevation environments. Five indicator species were chosen to assess 27 potential changes in each climate zones (see table). 28 Presence or absence of the indicator species were compared between baseline records and 29 20-year re-survey data for each site, providing a tally of indicator species gains, losses, and net 30 change in detection. National Forest System lands on the west slope of the Cascades Range are 31 mostly comprised of montane and higher elevation climate zones. Thus, a net increase in low 32 elevation, “warm low-elevation climate species,” such as those more common to the Willamette 33 Valley, could suggest warming. Conversely, a net increase in “cold climate high-elevation 34 species” might suggest cooling. 35 Analysis of net change in lichen species per climate zone showed an increase in three 36 “warm species” and all “cool species.” Two “cold species” increased, and two decreased (see 37 table). This suggests that some low- and mid-elevation species may have moved upslope during 38 the past 20 years. A few high-elevation “cold species” increased, perhaps because they have 39 broader climatic tolerances. Even within indicator groups, some species might have greater 40 tolerances to some environmental conditions. For example, Parmeliopsis hyperoptera can 41 tolerate a wider range of winter temperatures than Cetraria merrillii. 42 These data will be re-analyzed when more data come available. Additional data and 43 analyses can be found at the Pacific Northwest Region section of the U.S. Forest Service Lichens 44 and Air Quality website: http://gis.nacse.org/lichenair/?page=reports#R6. 45 DRAFT 2

46 Box 4.2—Use of departure metrics in assessing landscape sustainability and restoration 47 needs 48 49 Departure from the natural range of variability (NRV) can be used to assess landscape resilience 50 and sustainability with respect to ecological disturbances and vegetation structure (Hann et al. 51 2003, Hessburg et al. 2005, Keane et al. 2009, Landres et al. 1999). Fire regime condition class 52 (FRCC) is a simple similarity matrix that compares the current terrestrial condition of a 53 landscape with the estimated NRV to assess departure (Barrett et al. 2010, Schmidt et al. 2002). 54 The natural range is typically based on modeled conditions prior to European settlement (1850 in 55 the Pacific Northwest). Area in each of five seral stages for each potential vegetation type in 56 each watershed is modeled. Potential vegetation is used because it provides a framework to 57 define fire regimes. Fire frequency and severity are also included in FRCC, but because 58 relatively few data exist for these attributes, fire frequency and severity are not included in 59 FRCC assessments in a mapped context (see the FRCC guidebook [Barrett et al. 2010] and 60 FRCC mapping tool guidebook [Jones and Ryan 2012]). 61 Over the past decade, a number of FRCC assessments have been conducted in the Pacific 62 Northwest (e.g., DeMeo et al. 2012, Macdonald et al. 2006), leading to the identification of 63 specific needs for management treatments by seral stage, potential vegetation type, and 5th field 64 hydrologic unit of capability (watershed) (Haugo et al. 2015). Needs were categorized as 65 disturbance (e.g., thinning, prescribed fire, or wildfire), succession (growth over time, including 66 maintenance prescribed burning), and disturbance followed by succession. If implemented, 67 treatments are expected to move watersheds toward a more sustainable range of variation that 68 would be resilient to uncharacteristic wildfire, insect and disease outbreaks, and other 69 disturbances. The goal of treatment is not to impede natural disturbances, but to facilitate their 70 operating in a more sustainable way over time. 71 Haugo et al. (2015) did not assess Washington forests west of the crest of the Cascade 72 Range, although a subsequent assessment is in preparation that will cover this area. Data from 73 this work can be used to identify forest structural restoration needs for a large number of 74 watersheds with the highest value for protection and restoration (see details in USDA FS 2008). 75 Although forest structural restoration needs are a useful summary metric for landscape 76 restoration needs, other terrestrial, aquatic, and socio-economic attributes must also be 77 considered in restoration design and implementation. DRAFT 3

4

78 Box 5.1—Fishers: reintroduction and climate change in Southwest Washington 79 80 The fisher (Martes pennanti Erxleben) is a medium-sized member of the mustelid family 81 whose historical range in the Pacific Northwest once extended through most of the 82 Cascade and Coast Ranges, including the Olympic Peninsula (Lewis and Stinson 1998, 83 Lofroth et al. 2010, Powell 1993). Fishers were extirpated from Washington in the mid- 84 1900s by trapping and habitat fragmentation, and only a few populations remained in 85 other parts of the western United States (Aubry and Lewis 2003, Ingram 1973). Trapping 86 of fishers in Washington was prohibited in 1933, but populations had already decreased 87 significantly. Comprehensive multi-agency surveys in the latter part of the 20th century 88 showed no fishers in the Washington Cascade Range (Aubry and Houston 1992, Aubry 89 and Lewis 2003, Lewis and Hayes 2004). Recent efforts to reintroduce the fisher to parts 90 of its historical range include releases in the Olympic Peninsula (2008-2010), Sierra 91 Nevada (2009-2011), and southern Washington Cascades (2015) (Lewis et al. 2012). 92 Fishers are dietary generalists, which enables them to shift their diets depending 93 on location, season, and prey abundance, although they require specific habitat conditions 94 (Zielinski and Duncan 2004). Fishers use large-diameter trees for denning and resting, so 95 old growth (and mature) forest stands are particularly important for maintaining healthy 96 populations (Jacobson et al. 2003, Powell 1993). A dense forest canopy ensures the 97 availability of adequate resting sites, habitat for prey species, and refuge from predators 98 (Buchanan et al. 2013, Carroll et al. 1999, Purcell et al. 2009, Zielinski et al. 2004). A 99 dense canopy also decreases snow depth on the ground, which is important for fishers 100 because they do not have subnivean mobility (Aubry and Houston 1992, Carr et al. 2007, 101 Krohn et al. 1995). Fishers prefer low- and mid-elevation forests and tend to avoid high 102 elevations because deep snow, lowered abundance of prey, dispersed tree cover, lack of 103 large trees, and a lower abundance of snags and downed wood make habitat conditions 104 less favorable (Davis et al. 2007, Jacobson et al. 2003, Powell 1993, Spencer et al. 2011). 105 The effects of climate change on fishers will depend on dispersal abilities, 106 population numbers, current and future habitat fragmentation, forest management, 107 climate-induced ecosystem changes, and shifting habitats of predators and prey (Lawler 108 et al. 2012). Distribution of quality habitat patches and the ability of fishers to travel 109 between these areas will be limiting factors (Olson et al. 2014). Protecting quality fisher 110 habitat and maintaining suitable connectivity will help ensure that fishers are able to 111 persist in a warmer climate. These types of populations are also at increased risk of 112 extirpation from genetic isolation. Ensuring that fisher populations in Southwest 113 Washington are large enough to withstand pressures from climate change will be an 114 important conservation strategy, because local habitat shifts or disturbances can 115 disproportionately affect small or isolated populations. 116 DRAFT 4

5

117 Box 5.2—South Prairie: climate change effects on a unique habitat 118 119 South Prairie meadow is a botanical special interest area in Gifford Pinchot National Forest. The 120 main meadow covers 35 ha, and small surrounding meadows contain additional area. Spring 121 snowmelt accumulates in the basin of South Prairie, creating a seasonal lake. Specific 122 mechanisms that cause the seasonal inundation of the meadow are not clearly understood. It is 123 possible that ice accumulation in the adjacent Big Lava Bed prevents drainage until weather 124 warms sufficiently to melt the ice. 125 A small grove of quaking aspen exists at the edge of the coniferous forested area that is 126 annually inundated. Large-diameter black cottonwoods (Populus trichocarpa Torr. & A. Gray ex 127 Hook.) are found within the meadow itself; these cottonwoods stand in the middle of the lake 128 when the meadow is flooded (see figure inset). South Prairie is also the type locality for pale 129 blue-eyed grass, first collected and described by Suksdorf in 1893, and hosts the largest and most 130 genetically variable known population of the species (Ruchty 2011). 131 Effects of climate change on South Prairie are difficult to project because numerous 132 factors interact to create this unique habitat. Warmer temperatures and reduced precipitation 133 falling as snow will likely affect seasonal flooding of the meadow by influencing potential ice 134 formation and timing of run-off. Species-specific phenologies will be affected by changes in 135 timing and amount of seasonal flooding. Earlier drying of the meadow could affect competition 136 from nonnative plants, encourage conifer encroachment, facilitate trampling and compaction 137 associated with increased recreational use, and increase susceptibility to wildfire. 138 139 Box 5.2—Large black cottonwood trees in South Prairie meadow (a) during summer, and (b) 140 during spring with seasonal flooding. (Photos by [a] U.S. Forest Service and [b] Brittanie Hanna- 141 Anderson) 142

143 a) b) 144 DRAFT 5

6

145 Box 6.1—The Recreation Opportunity Spectrum 146 147 The Recreation Opportunity Spectrum (ROS) is a classification tool that has been used by federal 148 resource managers since the 1970s to provide visitors with varying challenges and outdoor 149 experiences (Clark and Stankey 1979, USDA FS 1990). The ROS is primarily used by the U.S. 150 Forest Service, but other federal agencies also incorporate its principles into recreation 151 management to some extent. The ROS classifies lands into six management class categories, 152 defined by setting and the probable recreation experiences: 153 Management class categories Urban Rural Roaded natural Semi-primitive motorized Semi-primitive non-motorized Primitive 154 155 Setting characteristics that define ROS include: 156 • Physical: type of access, remoteness, size of the area 157 • Social: number of people encountered 158 • Managerial: visitor management, level of development, naturalness (evidence of visitor 159 impacts and/or management activities) 160 161 The ROS is helpful for determining the types of recreational opportunities that can be provided. 162 After a decision has been made about the opportunity desirable in an area, the ROS provides 163 guidance about appropriate planning approaches and standards by which each factor should be 164 managed. Decisionmaking criteria include: (1) the relative availability of different opportunities, 165 (2) their reproducibility, and (3) their spatial distribution. The ROS Primer and Field Guide 166 (USDA FS 1990) specifically addresses access, remoteness, naturalness, facilities and site 167 management, social encounters, and visitor impacts. The ROS can be used to: 168 • Inventory existing opportunities 169 • Analyze the effects of other resource activities 170 • Estimate the consequences of management decisions on planned opportunities 171 • Link user desires with recreation opportunities 172 • Identify complementary roles of all recreation suppliers 173 • Develop standards and guidelines for planned settings and monitoring activities 174 • Help design integrated project scenarios for implementing resource management plans 175 176 In summary, the ROS approach provides a framework that allows federal land managers to 177 classify recreational sites and opportunities and to allocate resources for improvements and 178 maintenance within the broader task of sustainable management of large landscapes. 179 DRAFT 6

7

180 Box 6.2—An Overview of Recreation Data for Gifford Pinchot National Forest 181 182 Based on the most recently available data (Fiscal Year 2011), a total of 1,071,000 people visit 183 Gifford Pinchot National Forest (including Mt. St. Helens National Volcanic Monument) each 184 year, including: 185 • 585,000 in day-use developed sites 186 • 383,000 in general forest areas 187 • 77,000 in overnight-use developed sites 188 • 26,000 in wilderness 189 190 Of these visitors, 191 • 66% are male, 34% are female 192 • 43% are between 40 and 59 years old 193 • 94% are Caucasian (99% of wilderness visitors are Caucasian) 194 • 67% visit 5 times or fewer per year 195 • 79% travel 80 km or more to reach a recreation site; 49% travel 160 km or more 196 • 33% come from households with an income of $50,000-$75,000; less than 25% come 197 from households with an income of $100,000-$150,000 198 • Average duration of a national forest visit is 14 hours, median duration is 5 hours; 199 average duration of overnight use in a developed site is 45 hours 200 • 72% are very satisfied with their visit, 20% are somewhat satisfied 201 • Household income for visitors is 202 o Under $25,000 – 5.8% 203 o $25,000-$49,999 – 14.1% 204 o $50,000-$74,999 – 33.3% 205 o $75,000-$99,999 – 16.5% 206 o $100,000-$149,999 – 22.6% 207 o Above $150,000 – 7.7% 208 209 Eleven activities account for nearly 87% of recreational activity: 210 • Viewing natural features – 26.8% 211 • Hiking/walking – 21.1% 212 • Hunting – 12.0% 213 • Driving for pleasure – 7.0% 214 • Relaxing – 5.7% 215 • Developed camping – 2.8% 216 • Cross-country skiing – 2.7% 217 • Fishing – 2.5% 218 • Viewing wildlife – 2.2% 219 • Picnicking – 2.0% 220 • Snowmobiling – 2.0% 221 222 Average totalDRAFT trip spending per party is $408; median spending per party is $50. 7

Box 7.1—Forest carbon principles

In an effort to integrate carbon management with planning processes and climate change responses, the U.S. Forest Service created six guiding principles for carbon stewardship in national forests. They are intended to assist with integration of carbon management in planning and implementation and with efforts to adapt forests to a changing climate. These are preliminary principles intended to be refined, updated, and validated based on field experience, emerging science, and higher level interpretation across Forest Service programs and authorities.

• Emphasize ecosystem function and resilience. Carbon sequestration capacity depends on sustaining and enhancing ecosystem function to maintain resilient forests adapted to changing climate and other conditions.

• Recognize carbon sequestration as one of many ecosystem services. Carbon sequestration is one of many benefits provided by forests, grasslands, and forest products, now and in the future. Carbon sequestration should be considered in context with other ecosystem services.

• Support a diversity of approaches in carbon exchange and markets. Recognize that decisions about carbon in Americas forests are influenced by ownership goals, policy, ecology, geography, socioeconomic concerns, and other factors.

• Consider system dynamics and scale in decision making. Evaluate carbon sequestration and cycling at large spatial and temporal scales. Explicitly consider uncertainties and assumptions in evaluating carbon sequestration consequences of forest and grassland management options.

• Use the best information and methods to make decisions about carbon management. Base forest management and policy decisions on the best available science-based knowledge and information about system response and carbon cycling in forests and wood products. Use this information wisely by dealing directly with uncertainties, risks, opportunities, and tradeoffs through sound and transparent risk management.

• Strive for program integration and balance. Carbon management is part of a balanced and comprehensive program of sustainable forest management and climate change response. As such, forest carbon strategies have ecological, economic, and social implications and interactions with other Forest Service programs and strategies, such as those for energy and water. DRAFT 8

Box 7.2—Invasive species, climate change, and ecosystem services

Climate change has the potential to alter ecological processes in ways that increase the societal and environmental impacts of invasive species (Pyke et al. 2008). A species is considered to be invasive if: (1) it is nonnative to the ecosystem under consideration, and (2) its introduction causes, or is likely to cause, economic or environmental harm or harm to human health (Executive Order 13112, 1999). As native plant communities are disrupted by changing climatic conditions, invasives may become more competitive, with subsequent cascading effects on biotic and abiotic components of ecosystems (Charles and Dukes 2007, Hellmann et al. 2008). Invasive species have broad climatic tolerances and large geographic ranges, and are effective at overcoming barriers to dispersal, tolerating changing environmental conditions, and acquiring resources (Pyke et al. 2008). As ecosystem structures and systems change, so do the processes and functions that sustain ecosystem services (Charles and Dukes 2007, Pejchar and Mooney 2009). The U.S. Forest Service National Strategic Framework for Invasive Species Management states that “exotic species invasions and variations in climate patterns represent two of the greatest challenges to maintaining the ecosystem services provided by natural systems” (USDA FS 2013). The Framework identifies several threats posed by invasives to ecosystem services including “clean water, recreational opportunities, sustained production of wood products, wildlife and grazing habitat, and human health and safety.” The Framework estimates that damage from invasive species worldwide is $1.4 trillion y-1, or 5 percent of the global economy (Pimental et al. 2001).

DRAFT 9