FOURTH NATIONAL CLIMATE ASSESSMENT
CH. 24: NORTHWEST

EXECUTIVE SUMMARY:
CH. 24: NORTHWEST

Summary Overview

Residents of the Northwest list the inherent qualities of the natural environment among the top reasons to live in the region. The region is known for clean air, abundant water, low-cost hydroelectric power, vast forests, extensive farmlands, and outdoor recreation that includes hiking, boating, fishing, hunting, and skiing. Warming and related climate changes are already affecting aspects of the Northwest’s identity, such as its natural resource economy and its cultural heritage that is deeply embedded within the natural environment. The built systems that support Northwest residents and the health of residents themselves are also already experiencing the effects of climate change. The communities on the front lines of climate change experience the first, and often the worst, effects. Frontline communities in the Northwest include Tribal and Indigenous peoples, the economically disadvantaged, and those most dependent on natural resources for their livelihoods.

The region has warmed substantially—nearly 2°F since 1900—and this warming is partially attributable to human-caused emissions of greenhouse gases (Abatzoglou et al 2014, Vose et al. 2017). Warmer winters have led to reductions in the mountain snowpack (Mote et al 2017, EPA 2016) that has historically blanketed the region’s mountains, reducing wildfire risk (Westerling et al 2006, Littell et al 2016) and provided a slow release of water for communities, agriculture, rivers, and soils. In 2015, record winter warmth led to record-low snowpack in much of the Northwest’s mountains as winter precipitation fell as rain instead of snow (Mote et al 2016), resulting in drought, water scarcity, and large wildfires that negatively affected farmers, hydropower, drinking water, salmon, and recreation. In addition, warmer ocean temperatures led to shifts in the marine ecosystem, challenges for salmon, and a large harmful algal bloom. The extreme climate events of 2015 have prompted Northwest states, cities, Tribes, and others to increase and prioritize climate preparedness efforts.

Extreme climate events, such as those that occurred during 2015, can provide a preview of what may be more commonplace under a warmer future climate. The Northwest will continue to warm during all seasons under all future emissions scenarios, although the rate of warming will depend on current and future emissions (Rupp et al 2016). The warming trend may be accentuated in certain mountain areas in late winter and spring (Mote et al 2016), further exacerbating snowpack loss and increasing the risk for insect infestations and wildfires. In central Idaho and eastern Oregon and Washington, vast mountain areas have already been transformed by mountain pine beetle infestations, wildfires, or both, but the western Cascades and coastal mountain ranges have less experience with these growing threats.

Fig. 24.2: Multiple Climate Stressors Affect Vulnerable Infrastructure

Figure 24.2: Extreme events such as floods, heat waves, wildfires, landslides, and drought play an important role in the vulnerability of infrastructure. The example above from Seattle City Light's Vulnerability Plan (Raymond 2015) illustrates how the utility’s assets, operations, and management goals are affected by a broad range of climate impacts and extreme events. Adaptation strategies to increase the resilience of the energy system must focus on multiple potential risks as well as environmental considerations. Source: Raymond 2015.

SHRINK

Average winter precipitation is expected to increase over the long term, but year-to-year variability in precipitation will also increase (Rupp et al 2016). Years of abnormally low precipitation and extended drought conditions are expected to occur throughout the century (Rupp et al 2016), and extreme events, like heavy rainfall and heat waves, are also anticipated to occur more often (Brewer and Mass 2016). Along the coast, severe winter storms may also occur more often, such as occurred in 2015 during one of the strongest El Niño events on record (Paek et al 2017), contributing to storm surge, large waves, coastal erosion, and flooding in low-lying coastal areas (Barnard et al 2017). Changes in the ocean environment, such as warmer waters, altered chemistry, sea level rise, and shifts in the marine ecosystems are also expected. These projected changes will affect the Northwest’s natural resource economy, cultural heritage, built infrastructure, recreation, and the health and welfare of Northwest residents.

CHAPTER 24:
NORTHWEST

Background

Coordinating Lead Author:
Charles Luce, United States Department of Agriculture, Forest Service
Chapter Lead:
Kris May, Silvestrum Climate Associates
Chapter Authors:
Joe Casola, Climate Impacts Group, University of Washington
Michael Chang, Makah Tribe
Jennifer Cuhaciyan, United States Bureau of Reclamation
Meghan Dalton, Oregon State University
Scott Lowe, Boise State University
Gary Morishima, Quinault Indian Nation
Philip Mote, Oregon State University
Alexander (Sascha) Petersen, Adaptation International
Gabrielle Roesch-McNally, United States Forest Service
Emily York, Oregon Health Authority
Review Editor:
Beatrice Van Horne, Northwest Climate Hub
USGCRP Coordinators:
Natalie Bennett, Program Assistant
Christopher Avery, Senior Manager
Susan Aragon-Long, Senior Scientist
some fancy citation with bold and italic…

Residents of the Northwest list the inherent qualities of the natural environment among the top reasons to live in the region. The Northwest is known for clean air, abundant water, low-cost hydroelectric power, vast forests, extensive farmlands, and an array of outdoor recreation that includes hiking, boating, fishing, hunting, and skiing. Warming and related climate changes are already affecting aspects of the Northwest’s identity such as its natural resource economy and its cultural heritage that is deeply embedded within the natural environment. The built systems that support Northwest residents and the health of residents themselves are also already experiencing the effects of climate change. The communities on the front lines of climate change experience the first, and often the worst, effects. Frontline communities in the Northwest include Tribal and Indigenous peoples, the economically disadvantaged, and those most dependent on natural resources for their livelihoods.

The region has warmed substantially—nearly 2°F since 1900—and this warming is partially attributable to human-caused emissions of greenhouse gases (Abatzoglou et al 2014, Vose et al. 2017). Warmer winters have led to reductions in mountain snowpack (Mote et al 2017, EPA 2016), which has historically blanketed the region’s mountains, reducing wildfire risk (Westerling et al 2006, Littell et al 2016) and providing a slow-release of water for communities, agriculture, rivers, and soils. In 2015, record winter warmth led to record-low snowpack in much of the Northwest’s mountains as winter precipitation fell as rain instead of snow (Mote et al 2016), resulting in drought, water scarcity, and large wildfires that negatively affected farmers, hydropower, drinking water, salmon, and recreation. In addition, warmer ocean temperatures led to shifts in the marine ecosystem, challenges for salmon, and a large harmful algal bloom.

Extreme climate events, such as those that occurred during 2015, can provide a preview of what may be more commonplace under a warmer future climate. The Northwest will continue to warm during all seasons under all future emissions scenarios, although the rate of warming will depend on current and future emissions (Rupp et al 2016). The warming trend may be accentuated in certain mountain areas in late winter and spring (Mote et al 2016), further exacerbating snowpack loss and increasing the risk for insect infestations and wildfires. In central Idaho and eastern Oregon and Washington, vast mountain areas have already been transformed by mountain pine beetle infestations, wildfires, or both, but the western Cascades and coastal mountain ranges have less experience with these growing threats.

Average winter precipitation is expected to increase over the long term, but year-to-year variability in precipitation will also increase (Rupp et al 2016). Years of abnormally low precipitation and extended drought conditions are expected to occur throughout the century (Rupp et al 2016), and extreme events, like heavy rainfall and heat waves, are also anticipated to occur more often (Brewer and Mass 2016). Along the coast, severe winter storms may also occur more often, such as occurred in 2015 during one of the strongest El Niño events on record (Paek et al 2017), contributing to storm surge, large waves, coastal erosion, and flooding in low-lying coastal areas (Barnard et al 2017). Changes in the ocean environment, such as warmer waters, altered chemistry, sea level rise, and shifts in the marine ecosystems are also expected. These projected changes will affect the Northwest’s natural resource economy, cultural heritage, built infrastructure, recreation, and the health and welfare of Northwest residents.

Natural Resource Economy

Climate change is already affecting the Northwest’s diverse natural resources, which support sustainable livelihoods and provide a robust foundation for Tribal and rural communities. Increasing temperatures, changing precipitation patterns, and changes in coastal ocean waters have already reduced agricultural and fishery productivity, while also providing new business opportunities for parts of the natural resource economy. Climate change is expected to continue affecting the natural resource sector, valued at over $180 billion per year, but the economic consequences will depend on future market dynamics and adaptation efforts. Proactive management can increase the resilience of natural resources and economies.

Linkage between Observed Climate and Regional Risks

The Northwest provides for a diverse natural resource economy, from coastal fisheries, to Douglas fir plantations, to vineyards, to semi-arid rangelands, to dryland and irrigated farms. The region is the Nation’s top producer of 28 agricultural products, one of the leading national producers of timber products, and is widely recognized for salmon and shellfish fisheries. The agriculture, forestry, and fisheries sectors accounted for over 700,000 jobs and more than $139 billion in sales in 2015 (Figure 24.1, Sorte et al 2016).

Fig. 24.1: Natural Resource Industry Jobs and Sales Revenues

Figure 24.1: Natural resources are a key part of the Northwest economy. As the climate changes, jobs and sales revenues in the natural resource sector are at risk. Jobs and sales figures presented are based on 2015 data for the natural resource sector in Idaho, Oregon, and Washington (Sorte et al. 2016). Source: U.S. Forest Service and Boise State University.

EXPAND

The Northwest’s diverse natural environments attract a wide array of outdoor enthusiasts. The Outdoor Industry Association (2012) estimates that the region’s outdoor recreation economy generates $41.6 billion in consumer spending each year and provides around 444,800 jobs. These economic benefits are particularly important in rural and Tribal communities whose income base is largely dependent on natural resource economies and the supporting industries. Outdoor activities, including skiing, boating, rafting, hunting, fishing, hiking, and backpacking, can be impacted by climate variability, whether through less summer water, warmer streams, later snowfall, or loss of forests. Comparing high-snowfall to low-snowfall years in the Northwest between 1999 and 2009, the low-snowfall years resulted in 2,115 fewer employees and $172.7 million in reduced ski resort revenues (Burakowski and Magnusson 2010). Impacts on the skiing industry were especially prominent during the warm 2015 winter when snowpack was at record lows (see 2015 Case Study).

Both the natural resource commodity sector and outdoor recreation industry are sensitive to short- and long-term climate variability. The record-setting 2015 drought and above average temperatures were a challenge for agriculture. The reduced availability of water for irrigation coupled with heat stress impacted production and livestock health (see 2015 Case Study). In Northwest forests, wildfires, insect, and disease-driven mortality have been more prevalent over the last two decades due to drought conditions and increased temperatures, and timber managers are adjusting to increased risk of loss by shortening rotation rates, reducing investment, and changing planted species.

Commercial fisheries are also sensitive to climate variability. River temperatures increase during warm and dry years, resulting in fish kills of migrating and spawning salmon that have consequences several years in the future. In 2015, water temperatures in the lower Columbia River and tributaries were higher than at any other time on record, leading to a high rate of fish mortality (NOAA Fisheries 2016). The record temperatures in 2015 were part of a long-term trend of declining low flows (Kormos et al 2016) and warming streams (Isaak et al 2012). Increasing ocean temperatures and acidity also impact fish survival, species abundance, and predator–prey distribution and timing (Bakun et al 2015). In 2015, the increased ocean temperatures were part of an ocean heat wave coined the “Blob,” which fueled a coast-wide harmful algal bloom that impacted commercial, recreation, and tribal subsistence fisheries (see 2015 Case Study, Bond et al 2015, and Ch. 9: Oceans & Marine Resources).

Future Climate Change Relevant to Regional Risks

Shifts in timing of water supply, such as earlier snowmelt and declining summer flows, can adversely impact irrigated crop productivity, particularly where access to reservoir water storage and/or groundwater is limited. Planning studies for Northwest reservoirs suggest a significant increased need for reservoir storage to meet future summer irrigation demands under climate change scenarios (Reclamation 2011, Reclamation 2016b). Irrigation demands among farmers in the Columbia River Basin are projected to increase 5% in response to climate change by the 2030s; however, actual water demands will vary depending on adaptive management decisions and crop requirements (Rajagopalan et al in review). For dryland wheat production, rising temperatures coupled with increased atmospheric CO2 and associated increases in plant water use efficiency may lead to improved wheat yields under both higher (RCP8.5) and lower (RCP4.5) scenarios through the end of the century (Karimi et al 2017, Stockle et al 2017).

Specialty crops, including apples and other tree fruits, are already experiencing changes. Earlier high spring temperatures have led to earlier flowering, which can lead to a mismatch with the availability of pollinators required for fruit setting (process of flowers becoming fruit) (Houston et al. 2017) and can effect fruit quality as well as yield, additionally, summer heat stress is a concern, which can lead to sunburn scald on apples and softer berry crops that can be damaged in transport and harvest (Houston et al 2017) which can affect overall fruit quality and the price farmers receive for their product. Heat stress can also decrease livestock health which increases parasite abundance (Polley et al 2013), and projected warmer and drier summer seasons may also reduce forage quality and quantity (Izzauralde et al 2011)), with varied impacts across forage and rangeland types (Neibergs et al 2017). Impacts to the quality and quantity of forage will likely impact farmers’ economic viability as they may be required to pay for additional feed or wait longer for their livestock to put on weight which affects the total price they receive per animal.

Forests in the interior Northwest are changing rapidly because of increasing wildfire (Littell et al 2016), insect and disease damage (Kolb et al 2016, Ritokova et al 2016), and drought mortality (Luce et al 2016), attributed largely to a changing climate. These changes are expected to increase as temperatures increase (Peterson et al 2013). For forests that grow in areas with snowpack, the declining snowpack will worsen summer drought conditions, increasing vulnerability to drought caused by year-to-year precipitation variability (Vose et al 2016). Although some forests in the region may increase in productivity, others will decrease (Latta et al 2009; Ch. 6: Forests). Timber supplies from the drier eastern Northwest forests are the most affected by climate-related risks (Insley and Lei 2007), resulting in intermittent and unpredictable timber supplies and depressed timber prices (Sims 2011) in an already difficult global market. This could affect mill investments and the long-term viability of forestry as an economic activity, particularly in the more remote areas of the region where transportation costs to mills are high.

The impacts on Northwest fisheries associated with ocean warming, acidification, and harmful algal blooms are expected to increase. This could lead to extensive fisheries closures across all of the region’s coastal fisheries, with severe economic and cultural effects on commercial and subsistence shellfish industries. The warming ocean will also result in range shifts, with some species shifting as far north as the Bering Sea, yet these changes may also open up new fishing opportunities in the Northwest (Cheung et al 2015).

Projections for increased stream temperature indicate a 22% reduction in salmon habitat in Washington by 2090, leading to $3.3 billion in economic losses due to a reduction in salmon populations and decreases in cold-water angling opportunities under a high emissions future (A1F1, Niemi 2009). Freshwater trout are also sensitive to habitat connectivity and wildfire; therefore, land management practices may impact how trout respond to climate change (Rieman et al 2010). Overall, commercial fishing performance and abundance is expected to decline under future climate change scenarios (Sanford 2002, Ainsworth et al 2011, Weatherdon et al 2016), and it is likely that fishery ranges will change due to altered species distributions (Wenger et al 2011, Cheung et al 2015).

Decreases in low- and mid-elevation snowpack and accompanying decreases in summer streamflow will impact snow- and water-based recreation, such as downhill and cross-country skiing, snowmobiling, boating, rafting, and fishing. Research indicates that climate change could result in a 72% reduction in snow-based recreation revenue (−$300 million) and visits (−4.2 million visits) annually in the Northwest (Wobus et al 2017). Impacts to snowpack and, consequently, winter recreation will occur later in higher-elevation and colder mountains in southern Idaho (Luce et al 2014a).

Challenges, Opportunities, and Success Stories for Reducing Risk

Climate change will have both positive and negative effects on the natural resource sector. A shift in plant hardiness zones, or the ability of a given plant to thrive in a specific location, is expected, changing the suitability of growing certain crops (Parker and Abatzoglou 2016, McCarl et al 2016). Northwest wine producers may see the potential for growing higher-quality and higher-value wine grape varietals (Jones 2010). To take advantage of these opportunities, farmers may have to make costly changes and investments in new farming practices and territories in advance of projected climate changes (Houston et al 2017, Diffenbaugh et al 2011). While livestock producers in the region have an advantage over other U.S. regions where climate change impacts may be more severe (Mauger et al 2015), production costs are likely to increase because of supplemental feeding and watering requirements and the need for reducing livestock numbers in response to warmer and likely drier summers (Neibergs et al 2017). Adaptive approaches that build agroecosystem resilience will be required to maintain agricultural livelihoods (see Box 24.1).

CAPTION: Supplemental watering of livestock in Eastern Oregon during 2015 drought. Photo credit: Oregonian/OregonLive.

SHRINK

The prevalence of wildfires, insect infestations, disease epidemics, and drought-induced dieback of Northwest forests have heightened forestry managers’ awareness of potential climate change impacts. Forest management adaptation strategies are being developed (Halofsky and Peterson 2016, Halofsky et al 2017); including strategies that address drought-related risks, improve the reliability of forest transportation infrastructure, and protect forest-related ecosystem services (Peterson et al 2011). Vulnerability assessments and adaptation plans have been completed, or are in progress, for almost every National Forest and Park in the region (Adaptation Partners, 2017).

Many of the changes to the ocean environment cannot be adapted to or reduced. However, the fisheries industry has a history of implementing adaptation strategies to work within the limits of the natural environment to maintain species abundance to avoid extinction or increase harvests, such as limited fishing seasons, developing quota systems, and expanding aquaculture (see Ch. 9: Ocean & Marine Resources).

In response to climate variability, many ski resorts have diversified their recreational offerings to include warmer-weather opportunities such as mountain biking and hiking (Scott & McBoyle 2007, Shih et al 2009).

Despite the many strategies for reducing risks, adaptive capacity is not uniform across the natural resource economy. Given the heterogeneity across climatic and natural resource industries in the region, it is not likely that productivity gains and losses will be felt equally across the broad diversity in the region.

Emerging Issues

Climate drivers such as increased temperatures, CO2 fertilization, and precipitation changes will impact pest, disease, and weed pressures (Davis et al 2014, Eigenbrode et al 2013). Improved modeling of these stressors on yields and crop quality will enhance the understanding of climate change effects and inform adaptation options (Stockle et al 2017) and assist in addressing farmers’ concerns about future pest and pathogen impacts in the region (Morton et al 2017a/b). Water shortfalls are also likely to continue during drought periods despite adaptation efforts focused on water efficiency and reducing water usage. Western water law assigns a priority date to each right based on seniority, so junior (or more recent) water rights are more likely to be adversely affected under shortage conditions than those with senior water rights. More studies would enhance the understanding of which watersheds are at the greatest risk and what, if any, changes could address water limitations in the future.

Although much is being researched with respect to the effects of climate change on forests and associated ecosystem services, far less has been explored with respect to timber markets. Even then, most of the focus has been on changes in forest productivity overall (for example, Latta et al 2009) and less on the consequences of disturbance. Research is absent on the effects of potential increases in supply volatility and the consequences for investment and ultimately on harvest and milling jobs.

Ocean acidification poses a direct threat to shellfish and other calcifying species, including microscopic creatures with calcium and silicon shells that are at the base of the food web (Bednarsek and Ohman 2015). Shellfish farms in the Northwest were critical to the installation of an ocean monitoring system to track pH (that is, the acidity of the waters, where a lower pH value indicates more acidic conditions). Although calcium carbonate can be used to increase seawater pH in a hatchery setting (Barton et al 2015), the same approach cannot be used in the open ocean to prevent shell dissolution (Scigliano 2012). The broader food web consequences of decline in these species is an area of active research. As fishery ranges and distributions shift, predator–prey relationships may change, and the timing of the availability of food resources has been noted as a potential concern (Bakun et al 2015).

There is a great deal of uncertainty regarding impacts on the economic viability of natural resource-based economies in the region, particularly the degree to which individual sectors are integrated into global commodity markets, which will vary immensely and will be difficult to predict (EPA 2017).

Natural World / Cultural Heritage

Valued aspects of Northwest heritage and quality of life—the natural environment, wildlife, outdoor recreation, and Tribal cultures—will change with the climate. Increasing temperatures, reduced water availability, changing snow conditions, forest fires, habitat fragmentation, and other changes are endangering the well-being of a wide range of wildlife, threatening popular recreational activities and tribal subsistence and culture. For the Tribes, the health and vitality of the salmon runs is a direct indicator of the wider health of the region.

Linkage Between Observed Climate and Regional Risks

The intangible values and aspects of the Northwest’s natural environment—wildlife and habitat, outdoor recreation, and the heritage and quality of life of Tribal and Indigenous communities—are at risk in a changing climate.

CAPTION: First Salmon ceremony of the Lummi Tribe, Washington. Tribes in the Northwest typically honor the first salmon caught in the season through Tribal ceremonies. Photo credit: NW Treaty Tribes, 2012.

SHRINK

The Northwest’s native wildlife is impacted by climate variability and change directly through temperature shifts, water availability, and extreme events, and indirectly through loss or fragmentation of habitat (Bellard et al 2012). While climatic changes may not directly increase mortality, they can alter the balance among competing species or predator–prey relationships (for example, Wenger et al 2011). Three wildlife categories are of principal concern: already sensitive or endangered species, snow-dependent species, and game species. While the first two groups of animals are generally negatively impacted by most climate changes, some game species, such as deer and elk, may thrive as habitat changes. Game species are of concern not because of their sensitivity, but by their notable value for recreational hunting and as key cultural resources for Tribes.

Impacts to wildlife will also impact sacred First Foods, or foods that Tribes have historically cultivated for subsistence, economic, and ceremonial purposes (see photo). First Foods vary among Tribes, but often include berries, roots, water, fish, and local wildlife (Lynn et al 2013, Thornton et al 2015; Poe et al 2015). The loss or decline of First Foods can have cascading physical and mental health impacts for Tribal and Indigenous peoples (Norton-Smith et al 2016; McOliver et al 2015; see: Key Message 5).

Nearly half of the region’s adults participated in wildlife-related recreation in 2010 (National Survey 2011). As temperatures increase, the demand for warm-weather outdoor and water-based recreation increases, and visitation rates at local, state, and national parks increase (Fisichelli et al 2015, Buckley & Foushee 2011, Whitehead et al 2016). However, popular winter sports and snow-based recreational activities, such as downhill skiing, cross-country skiing, and snowmobiling, have been dramatically impacted by reduced snowfall. Low-snowfall years, which are becoming more common, result in far fewer skiing visits, meaning that residents and visitors are losing desirable skiing opportunities in warmer years. When summer streamflows and reservoir levels are lower, boating, and other water-based recreation opportunities also decline.

Future Climate Change relevant to Regional Risks

Wildlife responses are varied and complex. Some species, such as cavity nesting birds, may benefit from greater disturbance (Latif et al 2016, Saab et al 2014). Others, particularly snow-dependent species, may be unable to persist under climate change (McKelvey et al 2011).

Game species are expected to have diverse responses to climate change. Longer dry seasons and more pronounced droughts may reduce wetland habitat extent and duration, causing changes in waterfowl movement. Increased fire disturbance, on the other hand, may increase shrub cover, a preferred food for deer and elk (Keay et al 1980); reduced winter snowpack may increase food availability in winter; and warmer temperatures reduce winter stress, all of which would support higher deer and elk populations. Changes in disease and disease-carrying insects and pests may represent the primary risk and limitation in the future (Inkley et al 2013).

Temperature-sensitive bull trout and salmon and water-dependent species, such as amphibians, are among the most vulnerable because of increased habitat fragmentation (Case et al 2015, Isaak et al 2010, Rieman et al 2007). Increased frequency of extreme events such as flooding, debris flows, and landslides may also alter habitats and cause local extinctions of aquatic species.

Increased winter streamflow and decreased summer flow may threaten salmon spawning (Goode et al 2013). Many tribal salmon hatcheries and reintroduction efforts are therefore at risk (Cozzetto et al 2013). Projected increases in winter storm intensity can lead to higher river flows and increased sediment loading that can bury salmon eggs and reduce salmon survival (Cozzetto et al 2013). Rising stream temperatures, ocean acidification, and loss of nearshore and estuarine habitat also increases salmon mortality across all phases of the salmon life cycle (Crozier and Hutchings 2014).

Tribal shellfish beds are threatened by sea level rise, storm surge, and ocean acidification (Lynn et al 2013, Ekstrom et al 2015). Species may be moving out of traditional hunting, gathering, and fishing areas, which can impact resource access for many Tribes (Papiez 2009, Cozzetto et al 2013). Increasing wildfire frequency and intensity are changing foraging patterns for elk and deer, and increased prevalence of invasive species and disease may diminish both wildlife and foraging for traditional plants, berries, roots, and seeds (Voggesser et al 2013).

CAPTION: Razor clamming draws crowds on the coast of Washington. This popular recreation activity is expected to decline due to ocean acidification, harmful algal blooms, warmer temperatures, and habitat degradation. Credit: NOAA photo, courtesy Vera Trainer.

SHRINK

In the winter, continued decreases in lower-elevation snowpack will impact snow-based recreation. Less snowpack and earlier melting of snowpack will result in decreased water availability, impacting the quality, quantity, and availability of water-based recreational opportunities, such as boating, rafting, and fishing.

Increased wildfire occurrence will impact air quality and reduce the opportunity for and enjoyment of all outdoor recreation activities, such as camping, biking, hiking, youth sports, and hunting. Degraded air quality will also directly impact human health and quality of life (see Key Message 4).

Recreational ocean fishing opportunities are expected to decline under future climate change scenarios (Sanford 2002; Ainsworth et al 2011, Weatherdon et al 2016), and it is likely that fishery ranges will change (Cheung et al 2015). Recreational razor clamming on the coast is also expected to decline due to cumulative effects of ocean acidification, harmful algal blooms, higher temperatures, and habitat degradation (see photo; see Key Message 1).

Challenges, Opportunities, and Success Stories for Reducing Risk

Historical and projected changes in amenities affecting the quality of life the Northwest, such as wildlife, recreation opportunities, and edible plants, form a key challenge for managers of these resources. Informed management, however, can reduce the consequences to those who enjoy and value these resources. Sensitive and endangered species currently require special management considerations due to historical habitat changes and past species declines. Management of these species can substantially constrain land and water management options, and the protection of these species may become more difficult as suitable habitat is lost.

Game species are already managed. Further management of waterfowl habitat may be important to maintain past hunting levels. If deer and elk populations increase, the pressures they place on plant ecosystems (including riparian systems) may require management beyond traditional harvest levels.

Many Tribes have begun managing First Foods and other resources through climate change vulnerability assessments and climate change adaptation planning. Allowing Tribes to manage their resources in a self-determined and culturally-sensitive manner can increase each Tribe’s adaptive capacity to respond to climate change impacts on Tribal lands, foods, health, and cultures (see Box 24.2, Montag et al 2014, Lynn et al 2013, Colombi and Smith 2014). Sometimes, the cultural practice of harvesting and consuming First Foods may outweigh physical health implications, demonstrating how integral cultural practices are to Tribal and Indigenous health (Donatuto et al 2011). Tribes also increase their adaptive capacity through regional networks, such as the Columbia River Inter-Tribal Fish Commission, that support Tribal and Indigenous planning and management (see Key Message 5).

As fisheries become stressed due to climate change, additional management strategies may be required to maintain fish populations. Strategies that focus on habitat quality and quantity are likely to be the most successful. New ocean fishing opportunities may also be created as fishery ranges change (Cheung et al 2015).

Emerging Issues

Research is already active in identifying resilient habitats (for example, Luce et al 2014b, Isaak et al 2016) and the means for maintaining habitat resilience in the face of increasing climate and disturbance pressure (Hessburg et al 2015). Habitat modeling that includes projections of natural resource shifts, fragmentation, and identification of new wildlife corridors will be beneficial in supporting land and water management decisions that benefit people, recreation, and the Northwest’s varied wildlife. The institutional network of land, wildlife, and fishery management agencies, tribes, and non-government conservation organizations has already successfully reversed negative trends in many fish and wildlife populations caused by other human activities. These same groups are exploring methods to improve fish and wildlife resilience in a changing climate. It is likely that species that are not currently listed as endangered could become endangered over the next century, and eventual extinctions are likely yet challenging to predict (Pacifici et al 2015).

First Foods are an important aspect of Tribal and Indigenous health and well-being (Donatuto et al 2014). Since many First Foods will be impacted directly and indirectly by climate change, they can be used as indicators in Tribal health assessments and Tribal climate adaptation plans (Amberson et al 2016; Donatuto et al 2016). However, more research to refine these indicators would better support decision making (Biedenweg et al 2016; Donatuto et al 2016, see Box 24.2).

Social indicators link a decline in quality of life in the Northwest to loss of recreational opportunities due to climate change impacts (Klos et al 2015), but the causal links are not well understood. Additionally, future human migration and population increases may alter the relationship and nature of recreation in the Northwest (Binder and Jurjevich 2016). As the population increases, the demand for snow-based recreation may also increase. However, it is not clear how the limited availability of snow-based recreation (for example, a shorter ski season) in the Northwest over the long-term may influence interest in snow sports in contrast to alternatives.

Infrastructure

Existing water, transportation, and energy infrastructure already face challenges from flooding, landslides, drought, wildfire, and heat waves. Future climate change raises the risk for many of these extreme events, potentially compromising the reliability of water supplies, hydropower, and transportation across the region. Isolated communities and those with systems that lack redundancy are the most vulnerable. Adaptation strategies that address more than one sector, or are coupled with social and environmental co-benefits, can increase resilience.

Linkage Between Observed Climate and Regional Risks

Infrastructure plays a critical role in keeping the Northwest’s economy running smoothly. Roads, highways, railways, and ports facilitate the movement of people and goods within the region, and support valuable import and export markets. Powerlines and substations maintain the reliable supply of electricity to homes, businesses, schools, and hospitals. Dams and reservoirs manage streamflow to minimize flood risks, generate electricity, and provide water supply for irrigation and human consumption. Groundwater wells act as an important water source for agriculture and drinking supplies across much of the region. Levees and seawalls prevent damage to homes and property along rivers and the coast. Culverts manage water flows to protect roadways from flooding, and assist with fish passage, including migrating salmon. Stormwater and wastewater systems help minimize flooding, especially in urban areas, and are critical for maintaining water quality. However, most infrastructure is designed for a historical climate, and damage and disruptions caused by extreme events demonstrate existing infrastructure vulnerabilities that are likely to increase in a changing climate (see Ch. 11: Built Environment).

Services provided by infrastructure can be disrupted during extreme weather and climate events, illustrating the sensitivity of these systems to climate variability and climate change (see Box 24.3). For example, in Tillamook County, Oregon, in December 2015, heavy rainfall coupled with high coastal water levels caused flooding that led to culvert failures, road closures, and reduced access to health facilities (ODOT and OHA, 2016). The event highlighted the need to maintain detour routes that were valuable in reaching communities that could have become isolated.

Heavy rainfall can lead to slope instabilities and landslides, which can close roadways and railways. Along the Amtrak Cascades Corridor, multiple coastal bluff landslides have blocked the tracks and shut down rail service. Each landslide results in a minimum 48-hour moratorium on commuter rail service. Landslides during winter storms have also closed major Interstates, such as the December 2015 closure of eastbound Interstate-90 near Snoqualmie Pass, and the February 2017 closure of westbound Interstate-90 near Issaquah.

Wildfires can result in road and railway closures, reduced water quality in reservoirs, and impacts on the energy sector. The Goodell wildfire in August 2015 forced Seattle City Light to de-energize transmission lines around its Skagit Hydroelectric Project for several days (Raymond, 2015). The combined impact of damages and lost power production was nearly $3 million (SCL, 2016).

Drought conditions also present challenges for infrastructure, especially water supplies. In Washington, the Department of Ecology allocated almost $7 million in drought relief funds in 2015 (WA ECY, 2016a,b). Relief grants were used to provide backup or emergency water supplies for irrigation or human consumption where wells were failing or pumping capacity was inadequate. These small and typically rural systems are relatively more vulnerable to drought impacts when compared to larger urban systems.

Future Climate Change Relevant to Regional Risks

Climate change threatens to raise the risk of many extreme events that affect infrastructure in the Northwest. Available vulnerability assessments for infrastructure show the prominent role that future extremes play. Since much of the existing infrastructure was designed and is managed for an unchanging climate, changes in the frequency and intensity of flooding, drought, wildfire, and heat waves affect the reliability of water, transportation, and energy services.

Hydrologic change will be an important driver of future climate stress on infrastructure. As higher temperatures increase the proportion of cold season precipitation falling as rain rather than snow, higher streamflow will occur in many basins, raising flood risks (Hamlet et al 2013; Snover et al 2013; Mauger et al 2015; Dalton et al 2017). An increased risk of landslides is also expected, as more mixed rain and melting snow events occur in low- to mid-elevation mountains. Increases in the amount of precipitation falling in heavy rainfall events (including “atmospheric rivers”; Warner et al 2015) are anticipated to magnify these risks. Along the coast, sea level rise will increase flood risks in low-lying areas and will magnify the potential for erosion and infrastructure damage during extreme events with high storm surge and wave hazards.

Multiple Climate Stressors Affect Vulnerable Infrastructure

Figure 24.2: Extreme events such as floods, heat waves, wildfires, landslides, and drought play an important role in the vulnerability of infrastructure. The example above from Seattle City Light's Vulnerability Plan (Raymond 2015) illustrates how the utility’s assets, operations, and management goals are affected by a broad range of climate impacts and extreme events. Adaptation strategies to increase the resilience of the energy system must focus on multiple potential risks as well as environmental considerations. Source: Raymond 2015.

SHRINK

Spring and summer streamflows are anticipated to decline in basins that have historically relied on snowmelt, and low flow periods may be more prolonged and more severe. If observed declines in higher elevation precipitation continue (Luce et al 2013), this would exacerbate low streamflow conditions (Kormos et al 2016), resulting in decreased water supply and reservoir storage. Climate change can affect water quality as well. Higher air temperatures, lower streamflow, and decreases in rainfall are expected to raise summer stream temperatures, making it more difficult to meet water quality standards. In coastal areas, sea level rise may increase the risk of saltwater intrusion into groundwater supplies.

Challenges, Opportunities, and Success Stories for Reducing Risk

Anticipated future impacts on infrastructure create opportunities and add urgency for addressing existing environmental and social goals. For example, actions by the city of Boise, Idaho to improve water quality are likely to minimize some of the impacts associated with a warmer climate. In Boise, a phosphorous removal facility will reduce the amount of phosphorous entering rivers, thereby reducing the need for water treatment facility upgrades (Boise, 2016) and perhaps also reducing the risk of downstream algal blooms risks, which are anticipated to become more common in a warmer climate.

In Portland and Multnomah County, Oregon, considerations of social equity are woven into adaptation and are included in specific objectives of the 2030 Climate Change Preparation Strategy (Portland and Multnomah County, 2014). For many socially vulnerable populations, limited access to transportation, businesses, and other community resources can inhibit their coping capacity to climate impacts. Addressing these disparities can have the added benefit of bolstering resilience (see Key Message 5).

In many parts of the Northwest, the lack of redundancy within infrastructure systems may be an important factor in limiting adaptive capacity. Understanding the risks associated with these systems remains a challenge, as impacts could emerge directly from climate events, or from the interaction of non-climate and climate stressors (such as equipment failure making a water system more susceptible to subsequent drought). For example, in the Washington Department of Transportation’s vulnerability assessment, “lifeline” roadways that serve as the only means to access communities often emerged as highly vulnerable (WSDOT, 2011). Disruptions to these roadways could cut off communities, preventing supplies or first responders from arriving. In a similar vein, the Washington Department of Health is examining aspects of groundwater systems that may contribute to climate vulnerability. They have found that many groundwater systems are “single source” and lack any back-up supplies (see Figure 24.4). If supplies are disrupted, either by climate or non-climate stressors, surrounding communities may be forced to transport water to their area, or to relocate to a place with a more reliable supply.

An additional challenge in addressing risks to infrastructure is cost. Projects for replacing, retrofitting, or improving dams, reservoirs, pipelines, culverts, roadways, electrical transmission and distribution systems, and shoreline protection can have costs in the billions (for example, Wilhere et al 2017).

Groundwater Supply at Risk in Washington State

Figure 24.3: Map of single source, Group A groundwater systems in Washington State. Single source systems lack a backup supply, which may make it harder for users to cope with disruptions. Depth of each well is indicated by color. Shallower wells (red and yellow) may be at higher risk for drought impacts. However, the importance of depth varies from aquifer to aquifer. Data from the Sentry database, WA Department of Health, Office of Drinking Water.

SHRINK

Emerging Issues

Tectonic uplift along the Cascadia Subduction Zone has reduced the impact of sea level rise for the Northwest. The continental plate and sea level have both been rising, resulting in minimal relative sea level rise (the difference between the elevation of the continental crust and sea level). A magnitude 8 or greater earthquake along the Cascadia Subduction Zone could result in a large-scale drop in elevation of the continental plate, dramatically increasing relative sea level rise and exacerbating flooding in low-lying, coastal areas (NRC 2012). To address this combined risk, infrastructure managers are beginning to consolidate planning for sea level rise and seismic hazards, as well as tsunami risks which can also arise from a major earthquake event. Going forward, it could be useful to identify strategies that enhance community resilience and emergency response capacity to many types of hazards and potential disruptions.

Infrastructure management is traditionally oriented to protecting assets and services in place. However, in some locations and for some risks, it may be more efficient to remove or abandon infrastructure and find alternatives (for example, relocating communities and distributing water or energy systems). The knowledge and experience is just emerging to identify thresholds when such transformative decisions might be appropriate.

Health

The ability of regional social and healthcare systems to expand quickly beyond normal service levels will fall short if cascading or acute hazards occur, exacerbating existing socioeconomic disparities. In addition to an increased likelihood of acute hazards and epidemics, disruptions in local economies and food systems could result in more chronic health risks. Organizations and volunteers that make up the Northwest’s collective safety net are already stretched thin with current demands and will be further challenged by climate stressors. The potential health co-benefits of future climate mitigation investments could help to counterbalance these risks.

Linkage between Climate Change and Regional Risks

Over the last few decades, warmer and drier conditions during summer have contributed to longer fire seasons (Dalton 2017). Wildfire smoke can be severe, particularly in communities in the eastern Northwest (Idaho DEQ 2013). Smoke events during 2004–2009 were associated with a 7.2% increase in respiratory hospital admissions among adults over 65 in the western United States (Liu et al 2017). In Boise, Idaho, 7 of the last 10 years have included smoke levels considered “unhealthy for sensitive groups” (including children) for at least a week during the fire season (Idaho DEQ data summary report), causing some cancellation of school-related sports activities.

During extreme heat events in King County, Washington, heat-related hospital admissions were 2% higher and deaths 10% higher (Isaksen et al 2015, 2016), with an increased demand for emergency medical services for children, outdoor laborers, and the elderly (Calkins et al 2016). Oregon has also recorded spikes in heat-related emergency room visits (OHA Hazard Report).

In the last several years, the region has seen an increase in some infectious diseases. An increase in Lyme Disease cases is associated with rising temperatures and changing tick habitat (Beard et al 2016). The Washington Department of Health’s vector surveillance program has observed an earlier onset of West Nile Virus-carrying mosquitos likely associated with higher temperatures (WA-DOH, WNV report). Before 1999, cryptococcal infections were limited to the tropics, but Cryptococcus gatti, the species that causes these infections, is now established in Northwest soil, with 76 cases occurring in Oregon in 2015 (Bancroft 2017). The Oregon Health Authority recorded spikes in cases of Salmonella and E-coli during months with extreme heat in 2015 (Bancroft 2017). A large outbreak of Shigellosis occurred in late 2015 affecting a large number of homeless people in the Portland-Metro region; this outbreak was associated with unusually extreme precipitation (Hines 2017).

Changes in drought conditions and increased water temperatures have increased the potential for freshwater harmful algal blooms in recreational waters (Paerl 2009), although there is little capacity among state health departments to monitor and track harmful algal blooms. Toxins from marine harmful algal blooms can accumulate in shellfish, leading to illnesses for those who eat them (Bethel 2013). In 2015, during the largest harmful algal bloom ever observed off the West Coast from California to Alaska, high levels of domoic acid led to the closure of shellfish harvesting in much of the Northwest (Milstein 2015, 2015 Case Study).

Children and youth, in general, will experience cumulative physical and mental health effects (such as heat stress, trauma from injury, or displacement) of climate change over their lifetimes (Perera, 2017) due to increased exposure to extreme weather events and increased toxic exposures, such as ground-level ozone (“smog”) or increased contaminants in drinking water caused by flood events. Beginning at the fetal development stage, environmental exposures to air or water pollution can increase the risk of impaired brain development (Clifford, 2016), stillbirth (Siddika, 2016) and pre-term births (Sun, 2015; Peterson, 2015). Infants and children can be disproportionately affected by toxic exposures because they eat, drink, and breathe more in proportion to their body size (Heindel, 2016). Natural disasters as well as gradual changes (like changing landscapes and livelihoods) caused by climate stressors increase the risk of anxiety, depression, and post-traumatic stress disorder (PTSD) (Clayton et al, 2014). Evidence shows that exposure to both pollution and trauma early in life is detrimental to near-term health, and an increasing body of evidence suggests that early-childhood health status influences health and socioeconomic status later in life (Anda and Brown 2010; Currie 2014).

Future Climate Change Relevant to Regional Risks

More frequent wildfires and poor air quality are expected to increase respiratory illnesses in the decades to come. Airborne particulate levels from wildfires are projected to increase 160% by mid-century under a lower scenario (RCP4.5) (Liu et al 2016), creating a greater risk of smoke exposure through increasing frequency, length, and intensity of smoke events (Liu et al 2016).

Projected increases in ground-level ozone (smog) and airborne allergens (Fann et al 2016) can further complicate respiratory conditions. There is a well-documented link between exposure to air pollution and risk of heart attack, stroke, some types of cancer, and respiratory diseases (Cosselman 2015), all of which are leading causes of death in the Northwest (CDC 2016). The portion of each health condition attributed to air pollution is unknown, but the social and economic cost of these diseases is large. In Oregon, the medical costs associated with heart attacks in 2011 alone were over $1.1 billion, and those associated with stroke were $254 million (OHA 2013).

Increases in average and extreme temperatures will increase the number of heat-related deaths (Sarofim et al 2016). Mid-century climate in Portland, Oregon, under a mid-high scenario (RCP6.0) may result in 81–118 more heat-related deaths, although this figure does not account for future population growth or possible adaptations (Schwartz et al 2015).

Future extreme precipitation events could increase the risk of exposure to water-related illnesses as the runoff introduces contaminants and pathogens (such as Cryptosporidium, which causes gastrointestinal illness) into drinking water (Trtanj et al 2016). In the Puget Sound, under a medium emissions scenario (SRES A1B), local atmospheric heating of surface waters is projected to result in 30 more days per year that are favorable to algal blooms and an increased rate of bloom growth (Moore et al 2015).

Climate impacts could lead to income loss and an increase in the number of people experiencing food insecurity (see Key Message 1) (Haggerty 2014). As an example, in early 2016 a marine harmful algal bloom impacted the local Long Beach, Washington, economy, which is largely dependent on shellfish, tourism, and service industries. The local Food Bank recorded an almost 25% increase in the number of families requesting assistance in the six months that followed (OPFB 2017). Climate-driven hardships can also affect mental health, resulting in outcomes ranging from stress to suicide (Clayton 2017). Oregon, Washington, and Idaho are all ranked in the top 10 states with the highest prevalence of mental illness and lowest access to mental health care (Nguyen 2016). Tribal and Indigenous communities will also face multiple physical and mental health challenges, with impacts to subsistence and cultural resources (see Key Messages 2 and 5). Tracking climate stressors and training related to climate anxiety and post-disaster trauma is not widespread among the region’s health workforce (Doppelt 2016).

Challenges, Opportunities, and Success Stories for Reducing Risk

Existing environmental health risks will be exacerbated by projected climate conditions (Haggerty 2014), yet 98% of local health departments in Oregon reported having only partial-to-minimal ability to identify and address environmental health hazards (OHA 2016).

With funding from the Centers for Disease Control and Prevention, Oregon has been able to make some headway on assessing climate change vulnerabilities (Haggerty 2015) and recently released a statewide climate and health resilience plan (York 2017). Five local health jurisdictions in Oregon are some of the first in the country to complete local climate and health adaptation plans. Interventions to address community-identified priorities range from providing water testing for domestic well users in drought prone areas to quantifying the health co-benefits of proposed transportation investments.

The Washington Department of Health has also added a climate position to begin integrating climate considerations into the state’s public health system. Together, Northwest states have launched the Northwest Climate and Health Network for public health practitioners to share resources and best practices. Idaho, Oregon, and Washington all have syndromic surveillance systems that provide near real-time data from emergency room visits. This health data has the potential to be layered with climate and environmental data (such as temperature and air quality data), but such analysis has not been carried out on a broad scale.

Incorporating more health and wellness considerations into climate decision-making can increase a community’s overall resilience. For example, preserving the ecological functions of an area can also promote Tribal and Indigenous health, while investing in active transportation and green infrastructure can also improve air quality and increase physical activity (Younger 2008).

Emerging Issues

Communities with higher rates of morbidity and mortality often have less adaptive capacity and are more vulnerable to climate stressors (Weis 2016). Many people living in the Northwest already struggle to meet basic needs that could serve as protective factors and these numbers could increase. For example, roughly 1 in 5 children in the region live in a food-insecure household (Idaho Food Bank 2015; Oregon Food Bank 2016; Northwest Harvest 2015), and are already at higher risk of poor health outcomes like asthma and diabetes (Cook 2004). Both the states of Washington and Idaho have had some of the largest increases in homeless populations in the United States and in 2016, Oregon had the highest rate of unsheltered homeless families with children (USHUD, 2016). People lacking adequate shelter face increased climate risks (such as direct exposure to extreme heat or winter storms), while also having increased vulnerability (such as poorer health and less access to resources).

Displacement and increased migration to the Northwest could place increasing pressures on housing markets, infrastructure, and health and social service systems (Whitley Binder 2016). However, the role of climate as a driver for migration to the Northwest is speculative; current population forecasts do not yet account for climate factors (Saperstein 2015).

Public health leaders in the Northwest are working to modernize health systems to better respond to and prepare for complex and emerging health risks. Coordinated Care Organizations (CCOs) in Oregon, which serve as Medicaid insurance providers, are beginning to invest in certain climate protections for members. For example, covering the cost of air conditioning units for patients at risk of heat-related illnesses ensure patients can remain in their homes (Peden 2016). More studies are needed to fully account for the cost savings associated with these kinds of health-related services.

Frontline Communities

Communities on the front lines of climate change experience the first, and often the worst, effects. Frontline communities in the Northwest include Tribal and Indigenous peoples, the economically disadvantaged, and those most dependent on natural resources for their livelihoods. These communities generally prioritize basic needs, such as shelter, food, and transportation; frequently lack economic and political capital; and have fewer resources to prepare for and cope with climate disruptions. However, the social and cultural cohesion inherent in many of these communities provides a foundation for building community capacity and increasing resilience.

Linkage Between Observed Climate and Regional Risks

Because people care about the place they live, a focus on places highlights the local material and symbolic contexts in which people create their lives, and through which those lives derive meaning (Adger 2011; Cunsolo Willox et. al., 2012). This is true for communities across the Northwest whether or not they are on the frontline of dealing with climate change. While there are many types of frontline communities (those communities likely to experience climate impacts first and worst) in the region, this chapter highlights three sets of communities: Tribes, farmworkers, and low-income populations in urban environments.

The effects of climate variability and extreme events are not felt equally across communities in the Northwest. Frontline communities have higher exposures, are more sensitive, and are less able to adapt for a variety of reasons (Crimmins et al 2016, Gamble et al 2016, Haggerty et al 2014) including enhanced occupational exposure (Bethel and Harger 2014), dependence on natural and cultural resources (Norton-Smith et al 2016), fewer economic resources (Gamble et al 2016), other demographic factors (Cutter et al 2003, US DHHS 2013), and gender (Vinyeta et al 2015). In addition, frontline communities frequently must overcome cumulative exposures (McOliver et al 2015) and multigenerational trauma (McOliver et al 2015, Brave Heart et al 2011). It is the interconnected nature of legacy exposure, enhanced exposure, higher sensitivity, and less capability to adapt that intensifies a community’s climate vulnerability (Haggerty et al 2014, Morello-Frosch et al 2011, Morello-Frosch et al 2009). Climate risks can affect the health, well-being, and livelihoods of these communities by causing acute health risks, such as physical injury during severe weather (Gamble 2016, Clayton et al 2017), and indirectly through chronic impacts, such as food insecurity or mental health conditions like post-traumatic stress disorder (PTSD) (see Key Message 4).

Future Climate Change Relevant to Regional Risks

Frontline communities generally prioritize meeting existing basic needs, such as shelter, food, and transportation. While climate risks vary from community to community, neighborhood to neighborhood, and even person to person, for frontline communities, climate variability, climate change, and extreme climate events can exacerbate existing risks, further impacting their ability to meet basic needs (Dodgen et al 2016).

Northwest Tribes directly depend on natural resources, both on and off reservations, and are among the first to experience climate impacts. In the United States, the history of colonization, such as the Indian Removal Act of 1830, coupled with ongoing management barriers, such as the status of “domestic dependent sovereigns”, have led to many challenges for Tribal and Indigenous climate adaptation (McNeeley 2017, Norton-Smith 2016, Box 24.5). The loss or reduced availability of First Foods (Key Message 2) can have broad physical, cultural, and spiritual impacts, including diabetes, heart disease, mental health impacts, and loss of cultural identity (Gamble et al 2016, McOliver et al 2015). This may be coupled with mental health impacts associated with intergenerational and historical trauma, alcohol abuse, suicide, and others impacts (Gamble et al 2016, Key Message 2).

Farmworkers are vital to the region, yet they often earn very low wages, face discrimination and workplace hazards. Farmworkers and their families often deal with both chronic and acute health impacts because of the high cost of health care and physically demanding work environments. Overall, farmworkers, who are largely immigrant laborers from Mexico, Central America, and South America, face distinct challenges and are more vulnerable due to structural causes that can lead to exploitation, discrimination, and violence (Quesada et al 2011). Climate change may exacerbate these existing stressors. While the Northwest is not typically considered a high-risk area for heat-related illness, heat waves are increasing in frequency, intensity, and duration (Bumbaco et al 2013) and will make heat-related illness more common in the future. Farmworkers can be particularly vulnerable to heat-related illness due to occupational exposure (heavy exertion and working outdoors) (Bethel and Hargen 2014), yet they often do not seek health care because of high costs, language barriers, and fear of deportation (Hernandez 2016). Working conditions, as well as cooling and hydration practices, vary across the region (Bethel et al 2017).

Economically disadvantaged urban communities live in neighborhoods with the greatest exposure to climate and extreme weather events and are therefore disproportionately affected by climate stressors (Dodman and Satterthwaite 2008, Shi et al 2016). Urban heat islands, worsening air quality (WHO 2016), less access to transit, increasing demands for food and energy, and proximity to pollution sites can lead to injury, illness, and loss of life for the urban poor (Dodman and Satterthwaite 2008; Seattle 2016a, Key Message 4). For instance, in the Northwest, increased risk of heat-related illnesses and deaths have been associated with socioeconomic status, age, race, and occupation (for example, outdoor labor) (Isaksen et al 2015, Jackson et al 2010, Davis et al 2016).

Challenges, Opportunities, and Success Stories for Reducing Risk

Many frontline communities are taking actions that begin to address these challenges. Indigenous people and Pacific Northwest Tribes have demonstrated a high degree of resilience and have adapted to changing environmental and social conditions for thousands of years (Norton-Smith 2016). The strong social networks and connectivity, present in many Tribal and Indigenous communities, can reduce vulnerability to climate change (Balbus et al 2016). Efforts to enhance communication and strengthen network connections between tribes and their partners can be seen across the region.

Acknowledging the risk of heat-related illness for outdoor workers, the state of Washington issued rules requiring employers to make certain changes to job sites during the summer season (from May 1 through September 30). For temperatures above certain thresholds, the employer is required to provide at least one quart of water per employee per hour, relieve employees from duty if they are showing signs of heat-related illness, and provide training for employees and supervisors on heat-related illness (Washington Administrative Code 296-62-095).

Economically disadvantaged populations and communities of color often face multiple barriers to participating in public processes where decisions about future climate-related investments are made. Organizations representing these frontline communities have found some success prioritizing leadership development through workshops and trainings that enable new and emerging voices to be heard in more formal policy settings. Engagement has partly been made possible by providing transportation, childcare, meals, and accessibility, and by using a relational worldview and trauma-informed approach to community capacity-building. There has also been a concerted effort at the city and county policy level to explicitly acknowledge and address race and social inequities alongside environmental concerns (USDN 2017, Seattle 2017, Seattle 2016a, Seattle 2016b, Portland-Multnomah County 2016). Example actions include targeting investments in frontline communities and providing job training and employment opportunities that help limit displacement and enhance resilience (Portland-Multnomah County 2016).

Emerging Issues

There is an emerging understanding of the importance of not only prioritizing climate change preparedness efforts in frontline communities, but also involving and empowering these groups in the decision-making and implementation of climate change plans and actions.

The physical and psychological connections people have with natural resources are complex, and additional research would aid understanding of how changing climate conditions will affect not only those natural resources but also the people who depend on them. How intersecting vulnerabilities, driven by a confluence of climatic, social, and economic factors, may compound, and accelerate risks in frontline communities is not yet fully understood. Finally, additional research would help to measure and evaluate how supporting frontline communities in the implementation of community-identified strategies might improve outcomes and increase not only climate resilience but equity and economic vitality in the Northwest and across the country.

Extreme climate variability provides a preview of what may be commonplace in the future

Overall, 2015 temperatures were 3.8°F above the 20th century average, with winter temperatures 6.5°F above 20th century average (NOAA 2017). The warm 2015 winter temperatures are illustrative of conditions that may be considered “normal” by mid-century (higher scenario, RCP8.5) or late century (lower scenario, RCP4.5) (Rupp et al 2016).

Winter, spring, and summer precipitation during 2015 for the Northwest were below normal (as compared to the 1970–1999 average) by 13%, 23%, 38.5%, respectively (NOAA 2017). Precipitation from January to June 2015 was the 7th driest on record for the region (4.61 inches below the 20th century average) (NOAA 2017). In general, most climate models project increases in future Northwest winter and spring precipitation with decreases in the summer, although some models project increases and others decreases in each season (Rupp et al 2016). The 2015 summer precipitation deficits are similar to the largest decreases (−34%) in summer precipitation projected for end of the century (2070–2099) under a higher scenario (RCP8.5) (Rupp et al 2016).

Oregon and Washington 2015 snowpacks were the lowest on record at 89% and 70% below average, respectively (Mote et al 2016). These levels are more extreme than projected under the higher scenario (65% below average; RCP8.5) by end of century (Gergel et al 2017). However, with continued warming, this type of low snowpack drought is expected more often. For example, the 2015 extreme low snowpack conditions in the McKenzie River Basin (which sits largely in middle elevation of the Oregon Cascades) could occur on average about once every 12 years under 3.6°F (2.0°C) of warming (Sproles et al 2017). For each 1.8°F (1°C) of warming, peak snow-water equivalent in the Cascades is expected to decline 22%–30% (Cooper et al 2016).

What happened? How were systems tested? What vulnerabilities were highlighted?

Impacts from the 2015 “snow drought” were widespread, including irrigation shortages, agricultural losses, limited snow- and water-based recreation, drinking water quality concerns, hydropower shortages, and fish die offs from impaired stream water quality. Many farmers received a reduced allocation of water, and irrigation water rights holders had their water shut off early; senior water rights holders had their water shut off early for the first time ever (OWRD 2017). For example, Treasure Valley farmers in eastern Oregon received only a third of their normal irrigation water because the Owyhee reservoir received inadequate river inflows to fill the reservoir for the third year in a row (Stevenson, 2016).

Agricultural-related impacts of the drought were numerous, including damaged crops, reduced yields, altered livestock management, fewer planted crops, and land left idle (for example, 20% of farm acres in Treasure Valley, Oregon, were left idle) (OWRD 2017). Estimated agricultural economic losses were between $633 million and $773 million in Washington, including losses of $7.76 million in blueberries, $13.9 million in red raspberries, $500 million in a selection of 15 crops that make up more than three-quarters of Washington’s cultivated acreage, and $33.28 million in the dairy industry (McLain et al 2017).

Low-elevation ski areas struggled to open during the 2014–2015 season. Mount Ashland Ski Area in southwest Oregon did not open at all (Stevenson, 2016) and Hoodoo Ski Area had their shortest season in 77 years of operations after closing for the season in mid-January (Sproles et al 2017). Visitation at Detroit Lake, a reservoir in the Cascade foothills, decreased by 26% due to historically low water levels—70 feet (21 meters) below reservoir capacity in July—and unusable boat ramps (Wisler 2016, Sproles et al 2017).

Many significant fish kills occurred in Northwest streams due to a naturally occurring disease that thrived in the unusually high stream water temperatures (OWRD 2017, Fears 2015). And, for the first time ever, Oregon implemented a statewide daily fishing curtailment beginning in July 2015 to limit added stress on the fish from fishing (OWRD 2017).

The lack of snowpack in 2015 in concert with extreme spring and summer precipitation deficits led to the most severe wildfire season in the Northwest’s recorded history with more than 1.6 million acres burned across Oregon and Washington, incurring more than $560 million in fire suppression costs (Sexton et al 2016). In Oregon, the cost of large fires in 2015 was 344% of the 10-year average of large-fire costs (OWRD 2017). The wildfire season resulted in transmission shut downs for Seattle City Light during the Gooddell Fire (see Key Message 3) and infrastructure damage for Idaho Power Company following the Soda Fire (DOE 2015). Smoke from the wildfires caused significant air quality and health concerns from late July through September, particularly in eastern Oregon and Washington, Idaho, Colorado, and Canada (Creamean et al 2016, Jaffe and Zhang 2017).

The ocean heat wave referred to as the “Blob” was first detected off the Pacific coast in 2013, and by 2014 it spanned the coast from Alaska to California (Bond et al 2015). In 2015, the largest harmful algal bloom recorded on the West Coast was associated with the Blob. High levels of multiple toxins, including domoic acid and paralytic shellfish toxins, closed a wide range of commercial, recreational, and tribal fisheries, including salmon, shellfish, and Dungeness crab along the entire NW coast (Cavole et al 2016, Jacox et al 2016, Peterson et al 2016, McCabe et al 2016).

During the 2015–2016 extreme El Niño winter, wave energy along the West Coast was about 50% above normal (Barnard et al 2017). This resulted in several major storms hitting northwestern Oregon, bringing record-breaking rainfall, high winds, and high tides. Tillamook County in Oregon experienced a state of emergency that included major highway and road closures due to flooding, failed culverts, landslides, and sinkholes. Disruptions in transportation networks affected access to food, healthcare, and social services (See Key Message 2) (ODOT and OHA 2016).

Who is doing what to increase resilience? What success stories are there?

The conditions in 2015 tested the capacity of existing systems and provided insights into potential future adaptation priorities. Several actions to increase resilience have already begun across multiple levels of governance. For example, the Oregon Drought Task Force was created to “review the State’s existing drought response tools, identify potential gaps, and make recommendations on tools and information needed to ensure that the State is prepared to respond during a drought in the future” (HB 4113, 2016). Washington assessed the economic impact on agriculture and recommended developing a plan “to assist growers and plan for a future that will include increased incidence of severe weather events such as the 2015 drought” (McLain et al 2017).

At the onset of the drought, anticipated agricultural losses were much higher than what occurred because of actions at the federal and state levels, and actions implemented by the farmers themselves (McLain et al 2017). This highlights the adaptive capacity of the agricultural sector. However, as 2015 conditions become more regular as a result of climate change, some farms will struggle to stay solvent despite adaptation interventions (McLain et al 2017).

In the Yakima Basin, irrigators, conservation groups, and state and federal agencies worked together to replenish the diminished tributary flows to save the salmon runs and riparian habitat during the drought. Water from the Yakima River was redirected through farm irrigation canals to seven tributaries. Although this further reduced the farmers’ irrigation water, they agreed to continue rerouting water to sustain the fish (NOAA Fisheries 2015).

Over the past 20 years, Tillamook County has experienced 17 Presidentially Declared Disasters caused by flooding and severe weather events, including those in December 2015, and is projected to face more severe winter storms in the years to come. Continuing efforts to build resilience within the health and transportation sectors in response to flooding hazards will help the county weather future storms (ODOT and OHA 2016).

REFERENCES

Abatzoglou, John T., and Crystal A. Kolden. 2011. "Climate change in western US deserts: potential for increased wildfire and invasive annual grasses." Rangeland Ecology & Management 64.5: 471-478.

Abatzoglou JT, Rupp DE, Mote PW. 2014. Seasonal Climate Variability and Change in the Pacific Northwest of the United States. Journal of Climate 27(5): 2125–2142. DOI: 10.1175/JCLI-D-13- 00218.1.

Abatzoglou JT, Williams AP. 2016. Impact of anthropogenic climate change on wildfire across western US forests. Proceedings of the National Academy of Sciences 113(42): 11770–11775. DOI: 10.1073/pnas.1607171113.

Adaptation Partners. 2017. Science-management partnerships focused on climate change adaptation in the western United States. http://adaptationpartners.org/index.php

Adger, N., Barnett. J., Chapin III, F.S., Ellemor, H., 2011. This Must Be the Place: Underrepresentation of Identity and Meaning in Climate Change Decision-Making. Global Environmental Politics, Volume 11, Number 2 May 2011, pp 1-24. Published by The MIT Press.

Adger, W. 2003. Social capital, collective action, and adaptation to climate change. Economic Geography 79(4):387–404.

Ainsworth, C.H., Samhouri,J.F., Busch,D.S., Cheung,W.W.L., Dunne,J., and Okey, T.A. (2011). Potential impacts of climate change on Northeast Pacific marine food webs and fisheries. ICESJ.Mar.Sci. 68,1217–1229.doi: 10.1093/icesjms/fsr043

Amberson, S., Biedenweg, K., James, J., & Christie, P. (2016). “The Heartbeat of Our People”: Identifying and Measuring How Salmon Influences Quinault Tribal Well-Being. Society & Natural Resources, 1920(May), 1–16. http://doi.org/10.1080/08941920.2016.1180727

Anda R. F., Brown D. W. (2010). Adverse childhood experience & population health in Washington: The face of a chronic public health disaster. Retrieved from http://www.fpc.wa.gov/publications/ACEs%20in%20Washington.2009%20BRFSS.Final%20Report%207%207%202010.pdf.

Bakun, A., and coauthors. 2015. Anticipated Effects of Climate Change on Coastal Upwelling Ecosystems. Current Climate Change Reports 1(2):85-93.

Balbus, J., A. Crimmins, J.L. Gamble, D.R. Easterling, K.E. Kunkel, S. Saha, and M.C. Sarofim, 2016: Ch. 1: Introduction: Climate Change and Human Health. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. U.S. Global Change Research Program, Washington, DC, 25–42. http://dx.doi.org/10.7930/J0VX0DFW

Bancroft J, Byster L, 2017. Selected Reportable Communicable Disease Summary. Oregon Health Authority: Portland, OR.

Barnard PL, Hoover D, Hubbard DM, Snyder A, Ludka BC, Allan J, Kaminsky GM, Ruggiero P, Gallien TW, Gabel L, McCandless D, Weiner HM, Cohn N, Anderson DL, Serafin KA. 2017. Extreme oceanographic forcing and coastal response due to the 2015-2016 El Nino. Nature Communications 8: 14365. DOI: 10.1038/ncomms14365.

Barton, A., Waldbusser, G. G., Feely, R. A., Weisberg, S. B., Newton, J. A., Hales, B., ... & McLaughlin, K. (2015). Impacts of coastal acidification on the Pacific Northwest shellfish industry and adaptation strategies implemented in response. Oceanography, 28(2), 146-159. doi:10.5670/oceanog.2015.38

Bartos MD, Chester MV. 2015. Impacts of climate change on electric power supply in the Western United States. Nature Climate Change 5(8): 748–752. DOI: 10.1038/nclimate2648.

Beamish, R. J., McFarlane, G. A., & Thomson, R. E. (1999). Recent declines in the recreational catch of coho salmon Oncorhynchus kisutc ... Canadian Journal of Fisheries and Aquatic Sciences, 56(3), 506–515.

Beard CB, Eisen RJ, Barker CM, Garofalo JF, Hahn M, Hayden M, Monaghan AJ, Ogden NH, Schramm PJ. 2016. Ch. 5: Vectorborne Diseases. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. US Global Change Research Program: Washington, DC, 129–156.

Bednarsek, N., and M. D. Ohman. 2015. Changes in pteropod distributions and shell dissolution across a frontal system in the California Current System. Marine Ecology Progress Series 523:93-103.

Bednaršek, N., Feely, R. a, Reum, J. C. P., Peterson, B., Menkel, J., Alin, S. R., & Hales, B. (2014). Limacina helicina shell dissolution as an indicator of declining habitat suitability owing to ocean acidification in the California Current Ecosystem. Proceedings of the Royal Society, Biological Sciences, 281(1785), 20140123. http://doi.org/10.1098/rspb.2014.0123

Bellard, C., C. Bertelsmeier, P. Leadley, W. Thuiller, and F. Courchamp (2012), Impacts of climate change on the future of biodiversity, Ecology letters, 15(4), 365-377.

Bethel J, Ranzoni S, Capalbo S. 2013. Human Health: Impacts and Adaptation: Chapter 7. In: Dalton MM, Mote PW and Snover AK (eds) Climate Change in the Northwest: Implications for Our Landscapes, Waters, and Communities. Island Press: Washington, DC, 181–206.

Bethel, J., and Harger, R., 2014. Heat-Related Illness among Oregon Farmworkers. International Journal of Environmental Research and Public Health. Sep; 11(9): 9273-9285. Published online 2014 Sep 5. doi: 10.3390/ijerph110909273

Bethel, J., Spector, J., and Krenz, J., 2017. Hydration and cooling practices among farmworkers in Oregon and Washington. Journal of Agromedicine. Vol 0 (0) http://dx.doi.org/10.1080/1059924X.2017.1318100

Biedenweg, K., Stiles, K., & Wellman, K. (2016). A holistic framework for identifying human wellbeing indicators for marine policy. Marine Policy, 64, 31–37. http://doi.org/10.1016/j.marpol.2015.11.002

Biedenweg, K., Hanein, A., Nelson, K., Stiles, K., Wellman, K., Horowitz, J., & Vynne, S. (2014). Developing Human Wellbeing Indicators in the Puget Sound: Focusing on the Watershed Scale. Coastal Management, 42(4), (online). http://doi.org/10.1080/08920753.2014.923136

Binder, LW and Jurjevich JR, 2016. Winds of Change? Exploring Climate Change-Driven Migration and Related Impacts in the Pacific Northwest. Environmental Migration Portal.
https://environmentalmigration.iom.int/winds-change-exploring-climate-change-driven-migration-and-related-impacts-pacific-northwest

Bishop, C.P. and C. R. Shumway. 2009. The Economics of Dairy Anaerobic Digestion with Coproduct Marketing. Review of Agricultural Economics 31(3): 394-410.

Black, Geoff, Michail Fragkias, Zeynep Hansen, Don Holley, Rob Humphrey, Scott Lowe. (2016). “Economic Impact and Importance of Power Boating in Idaho: A County-Level Study”. Idaho Parks and Recreation. https://parksandrecreation.idaho.gov/sites/default/files/uploads/documents/IDPR%20Powerboating%20Report_December%202016.pdf

Bond NA, Cronin MF, Freeland H, Mantua N. 2015. Causes and impacts of the 2014 warm anomaly in the NE Pacific. Geophysical Research Letters 42(9): 2015GL063306. DOI: 10.1002/2015GL063306.

Boyte, Stephen P., Bruce K. Wylie, and Donald J. Major. "Cheatgrass Percent Cover Change: Comparing Recent Estimates to Climate Change− Driven Predictions in the Northern Great Basin." Rangeland Ecology & Management 69.4 (2016): 265-279.

Brave Heart, M.Y.H.; Chase, J.; Elkins, J.; Altschul, D.B. Historical trauma among Indigenous peoples of the Americas: Concepts, research, and clinical considerations. J. Psychoact. Drugs 2011, 43, 282–290.

Brewer MC, Mass CF. 2016. Projected changes in heat extremes and associated synoptic/mesoscale conditions over the northwest U.S. Journal of Climate. DOI: 10.1175/JCLI-D- 15-0641.1.

Bromirski PD, Flick RE, Miller AJ. 2017. Storm surge along the Pacific coast of North America. Journal of Geophysical Research: Oceans 122(1): 441–457. DOI: 10.1002/2016JC012178.

Brown, T. T., and D. R. Huggins. "Soil carbon sequestration in the dryland cropping region of the Pacific Northwest." Journal of Soil and Water Conservation 67.5 (2012): 406-415.

Buckley, L. B., & Foushee, M. S. (2011). Footprints of climate change in US national park visitation. International journal of biometeorology, 56(6), 1173-1177. DOI 10.1007/s00484-011-0508-4

Bumbaco, K., Dello, K., and Bond, N., 2013. History of Pacific Northwest Heat Waves: Synoptic Patterns and Trends. Journal of Applied Meteorology and Climatology. American Meteorological Society. 52 (1618-1631) https://doi.org/10.1175/JAMC-D-12-094.1

Bureau of Reclamation and Washington Department of Ecology (Reclamation and Ecology). 2011. Yakima River Basin Study: Volume 1 - Proposed Integrated Water Resource Management Plan. https://www.usbr.gov/watersmart/bsp/docs/finalreport/Yakima/YakimaRiverBasinStudy-ResourceMgmtPlan.pdf

Bureau of Reclamation (Reclamation). 2011. Climate and Hydrology Datasets for Use in the River Management Joint Operating Committee Climate Agencies’ Longer-Term Planning Studies: Part II – Reservoir Operations Assessment for Reclamation Tributary Basins. U.S. Department of the Interior, Bureau of Reclamation, Pacific Northwest Regional Office, Boise, Idaho. January 2011.

Bureau of Reclamation (Reclamation). 2016a. Columbia River Basin Impact Assessment. Prepared by U.S. Department of the Interior, Bureau of Reclamation, Pacific Northwest Region, Boise, Idaho. March 2016

Bureau of Reclamation (Reclamation). (2016b). SECURE Water Act Section 9503(c) – Reclamation Climate Change and Water. Prepared for United States Congress. Bureau of Reclamation, Policy and Administration: Denver, CO.

Bureau of Reclamation (Reclamation). 2015a. Infrastructure Investment Strategy. Prepared by U.S. Department of the Interior, Bureau of Reclamation, Denver, Colorado. May 2015.

Bureau of Reclamation (Reclamation). 2015b. Hood River Basin Study. Prepared by U.S. Department of the Interior, Bureau of Reclamation, Pacific Northwest Region, Boise, Idaho. November 2015. https://www.usbr.gov/watersmart/bsp/docs/finalreport/hoodriver/hoodriverbasinstudy.pdf

Bureau of Reclamation (Reclamation). 2015c. Henrys Fork Basin Study. Prepared by U.S. Department of the Interior, Bureau of Reclamation, Pacific Northwest Region, Boise, Idaho. January 2015. https://www.usbr.gov/watersmart/bsp/docs/finalreport/HenrysFork/HenrysForkBasinStudyReport.pdf

Burakowski, Elizabeth, and Matthew Magnusson. (2012). Climate Impacts on the Winter Tourism Economy in the United States. Natural Resources Defense Council and Protect Our Winters Report.

Burge, Colleen A., et al. "Climate change influences on marine infectious diseases: implications for management and society." Annual review of marine science 6 (2014): 249-277.

Calkins, M. M., Isaksen, T. B., Stubbs, B. A., Yost, M. G., & Fenske, R. A. (2016). Impacts of extreme heat on emergency medical service calls in King County, Washington, 2007–2012: relative risk and time series analyses of basic and advanced life support. Environmental Health, 15(1), 13. http://doi.org/10.1186/s12940-016-0109-0

Case, M. J., J. J. Lawler, and J. A. Tomasevic (2015), Relative sensitivity to climate change of species in northwestern North America, Biological Conservation, 187, 127-133.

Caulkin, MM, Isaksen TB, Stubbs BA, Yost MG, Fenske RA (2016) Impacts of extreme heat on emergency medical service calls in King County, Washington, 2007-2012: relative risk and time series analyses of basic and advanced life support. Environmental Health: Jan 28;15:13. doi: 10.1186/s12940-016-0109-0.

Cavole L-CM, Demko AM, Diner RE, Giddings A, Koester I, Pagniello CMLS, Paulsen M-L, Ramirez-Valdez A, Schwenck SM, Yen NK, Zill ME, Franks PJS. 2016. Biological Impacts of the 2013-2015 Warm-Water Anomaly in the Northeast Pacific. Oceanography 29(2): 273–285. DOI: 10.5670/oceanog.2016.32.

Centers for Disease Control and Prevention (CDC) (2016). National Center for Health Statistics: Stats of the States. Retrieved from: https://www.cdc.gov/nchs/pressroom/stats_of_the_states.htm

Chambers and Pellant. 2008. Climate Change Impacts on Northwestern and Intermountain United States Rangelands. Society for Rangeland Management.

Cheng, T. K., D. F. Hill, J. Beamer, and G. Garcia-Medina. 2015. Climate change impacts on wave and surge processes in a Pacific Northwest (USA) estuary. Journal of Geophysical Research-Oceans 120(1):182-200.

Cheung WWL, Brodeur RD, Okey TA, Pauly D. (2015). Projecting future changes in distributions of pelagic fish species of Northeast Pacific shelf seas. Progress in Oceanography 130: 19–31. DOI:10.1016/j.pocean.2014.09.003.

City of Boise, 2016. Boise opens National Precedent Setting Phosphorous Removal Facility. http://publicworks.cityofboise.org/news-releases/2016/08/boise-city-opens-national-precedent-setting-phosphorus-removal-facility/

City of Portland and Multnomah County, 2014. Climate Change Preparation Strategy. Risk and Vulnerability Assessment. https://multco.us/file/36550/download

Clayton, S., Manning, C. M., Krygsman, K., & Speiser, M. (2017). Mental Health and Our Changing Climate: Impacts, Implications, and Guidance. Washington, D.C.: American Psychological Association, and ecoAmerica.

Clayton, S., Manning, C. M., & Hodge C. (2014). Beyond storms & droughts: The psychological impacts of climate change. Washington, DC: American Psychological Association and ecoAmerica.

Clifford, A., Lang, L., Chen, R., Anstey, K., & Seaton, A. (2016). Exposure to air pollution and cognitive functioning across the life course--A systematic literature review. Environ Res., 383-98. doi:10.1016/j.envres.2016.01.018

Colombi, B. J. (2012). Salmon and the Adaptive Capacity of Nimiipuu (Nez Perce) Culture to Cope with Change. American Indian Quarterly, 36(1), 75–97. http://doi.org/10.5250/amerindiquar.36.1.0075

Colombi, Benedict J.; Smith, Courtland L. 2014. Insights on adaptive capacity: three indigenous Pacific Northwest historical narratives. Journal of Northwest Anthropology. 48(2): 189-201.

Columbia River Long Term Water Supply and Demand Forecast. 2016. Department of Ecology (Washington). http://www.ecy.wa.gov/programs/wr/cwp/2016Forecast.html

Columbia River Inter-Tribal Fish Commission (CRITFC). (2017). Salmon Culture. http://www.critfc.org/ salmon-culture/tribal-salmon-culture. Accessed 13 April 2017.

Cook J, et al (2004). Food Insecurity is Associated with Adverse Health Outcomes among Human Infants and Toddlers. The Journal of Nutrition, vol 134, no 6, 1432-1438.

Cooper MG, Nolin AW, Safeeq M. 2016. Testing the recent snow drought as an analog for climate warming sensitivity of Cascades snowpacks. Environmental Research Letters 11(8): 84009. DOI: 10.1088/1748-9326/11/8/084009.

Cosselman KE, Navas-Acien A, Kaufman JD (2015). Environmental factors in cardiovascular disease. Nature Reviews Cardiology 12, 627-642. doi:10.1038/nrcardio.2015.152

Coutts, C., Hahn, M. (2015). Green infrastructure, ecosystem services, and human health. Int. J. Environ. Res. Public Health. 12: 9768-9798. Doi: 10.3390/ijerph120809768

Cozzetto, K., K. Chief, K. Dittmer, M. Brubaker, R. Gough, K. Souza, F. Ettawageshik, S. Wotkyns, S. Opitz-Stapleton, and S. Duren (2013), Climate change impacts on the water resources of American Indians and Alaska Natives in the US, Climatic Change, 120(3), 569-584. http://doi.org/10.1007/s10584-013-0852-y

Creamean JM, Neiman PJ, Coleman T, Senff CJ, Kirgis G, Alvarez RJ, Yamamoto A. 2016. Colorado air quality impacted by long-range-transported aerosol: a set of case studies during the 2015 Pacific Northwest fires. Atmospheric Chemistry and Physics 16(18): 12329–12345. DOI: 10.5194/acp-16-12329-2016.

Crimmins A, Balbus J, Gamble JL, Beard CB, Bell JE, Dodgen D, Eisen RJ, Fann N, Hawkins MD, Herring SC, Jantarasami L, Mills DM, Saha S, Sarofim MC, Trtanj J, Ziska L. 2016. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. US Global Change Research Program: Washington, DC, 312. http://dx.doi.org/10.7930/J0R49NQX

Crozier, L. 2016. Impacts of Climate Change on Salmon of the Pacific Northwest: A review of scientific literature published in 2015. Literature review from: Fish Ecology Division of the Northwest Fisheries Science Center. Seattle, WA. Accessed: June 9 2017. https://www.nwfsc.noaa.gov/assets/4/9042_02102017_105951_Crozier.2016-BIOP-Lit-Rev-Salmon-Climate-Effects-2015.pdf.

Crozier, Lisa G., and Jeffrey A. Hutchings. "Plastic and evolutionary responses to climate change in fish." Evolutionary Applications 7.1 (2014): 68-87.

Cunsolo Wilcox, A., Harper, S., Ford, J., Landman, K., Houle, K., Edge, V., 2012. “From This Place and of This Place:” Climate Change, Sense of Place, and Health in Nunatsiavut, Canada. Soc Sci Med 75(3), 5380547 April 16. PubMed: 22595069 DOI: 10.1016/j.socscimed.2012.03.043

Currie J, Zvin JG, Mullins J, Neidell M. 2014. What do we know about short-and long-term effects of early-life exposure to pollution? Annual Review of Resource Economics, November 2014. https://doi.org/10.1146/annurev-resource-100913-012610

Cutter, S. L., Boruff, B. J. and Shirley, W. L. (2003), Social Vulnerability to Environmental Hazards*. Social Science Quarterly, 84: 242–261. doi:10.1111/1540-6237.8402002

Dalton, M.M., K.D. Dello, L. Hawkins, P.W. Mote, and D.E. Rupp (2017) The Third Oregon Climate Assessment Report, Oregon Climate Change Research Institute, College of Earth, Ocean and Atmospheric Sciences, Oregon State University, Corvallis, OR.

Dalton, M., P.W. Mote, and A.K. Snover, eds., 2013: Climate Change in the Northwest: Implications for Our Landscapes, Waters, and Communities. 224 pp. Island Press.

Daly EA, Brodeur RD, Auth TD. 2017. Anomalous ocean conditions in 2015: impacts on spring Chinook salmon and their prey field. Marine Ecology Progress Series 566: 169–182. DOI: 10.3354/meps12021.

Daly, E. A., & Brodeur, R. D. (2015). Warming Ocean Conditions Relate to Increased Trophic Requirements of Threatened and Endangered Salmon. PLoS ONE, 10(12), e0144066. http://doi.org/10.1371/journal.pone.0144066

Davis, R. E., Hondula, D. M., & Patel, A. P. (2016). Temperature observation time and type influence estimates of heat-related mortality in seven U.S. cities. Environmental Health Perspectives, 124(6), 795–804. http://doi.org/10.1289/ehp.1509946

Davis et al. 2014. Differing Contributions of density dependence and climate to the population dynamics of three eruptive herbivores. Ecological Entomology 39:566-577.

Delcour, I., Spanoghe, P., Uyttendaele, 2015, Literature review: Impact of climate change on pesticide use. Food Research International. Volume 68, pp 7-15, Elsevier. https://doi.org/10.1016/j.foodres.2014.09.030

Di Lorenzo E, Mantua N. 2016. Multi-year persistence of the 2014/15 North Pacific marine heatwave. Nature Climate Change 6: 1042–1047. DOI: 10.1038/nclimate3082.

Dietrich, Joseph P., et al. 2014. "The impact of temperature stress and pesticide exposure on mortality and disease susceptibility of endangered Pacific salmon." Chemosphere 108 (2014): 353-359.

Diffenbaugh, Noah S., et al. 2011. "Climate adaptation wedges: a case study of premium wine in the western United States." Environmental Research Letters 6.2 (2011): 024024.

Dittmer, K. (2013), Changing streamflow on Columbia basin tribal lands—climate change and salmon, Climatic Change, 120(3), 627-641.

Dodgen D, Donato D, Kelly N, La Greca A, Morganstein J, Reser J, Ruzek J, Schweitzer S, Shimamoto MM, Thigpen Tart K, Ursano R. 2016. Ch. 8: Mental Health and Well-Being. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. US Global Change Research Program: Washington, DC, 217–246.

Dodman, D., & Satterthwaite, D. (2008). Institutional Capacity, Climate Change Adaptation and the Urban Poor. IDS Bulletin, 39(4), 67–74. http://doi.org/10.1111/j.1759-5436.2008.tb00478.x

Donatuto, J. L., Satterfield, T. a., & Gregory, R. (2011). Poisoning the body to nourish the soul: Prioritising health risks and impacts in a Native American community. Health, Risk & Society, 13 (December 2014), 103–127. http://doi.org/10.1080/13698575.2011.556186

Donatuto, J., Grossman, E. E., Konovsky, J., Grossman, S., & Campbell, L. W. (2014). Indigenous Community Health and Climate Change: Integrating Biophysical and Social Science Indicators. Coastal Management, 42(4), 355–373. http://doi.org/10.1080/08920753.2014.923140

Doppelt, B. (2016). Transformational Resilience: How Building Human Resilience to Climate Disruption Can Safeguard Society and Increase Wellbeing. Greenleaf Publishing, UK.

Doubleday, A. J., & Hopcroft, R. R. (2015). Interannual patterns during spring and late summer of larvaceans and pteropods in the coastal Gulf of Alaska, and their relationship to pink salmon survival. Journal of Plankton Research, 37(1), 134–150. http://doi.org/10.1093/plankt/fbu092

Dulal HB, 2017. Making cities resilient to climate change: identifying “win-win” interventions. The International Journal of Justice and Sustainability, Vol. 22, Issue 1, pp 106-125. DOI: 10.1080/13549839.2016.1168790

Duran, B., Duran, E., & Yellow Horse Braveheart, M. (1998). Native Americans and the trauma of history. In R. Thornton (Ed.), Studying native America: Problems and prospects in Native American studies (pp. 60-76). Madison: University of Wisconsin Press. Google Scholar

Durkalec, A., Furgal, C., Skinner, M., & Sheldon, T. (2015). Climate change influences on environment as a determinant of Indigenous health: Relationships to place, sea ice, and health in an Inuit community. Social Science and Medicine, 136-137, 17–26.

Eigenbrode et al. 2013. Agriculture: Impacts, Adaptation and Mitigation in Climate Change in the Northwest. Washington, DC, Island Press.

Ekstrom, J. A., Suatoni, L., Cooley, S. R., Pendleton, L. H., Waldbusser, G. G., Cinner, J. E., … Portela, R. (2015). Vulnerability and adaptation of US shellfisheries to ocean acidification. Nature Climate Change, 5(3), 207–214. http://doi.org/10.1038/nclimate2508

EPA. 2017. Multi-Model Framework for Quantitative Sectoral Impacts Analysis: A Technical Report for the Fourth National Climate Assessment. U.S. Environmental Protection Agency, EPA 430-R-17-001.

EPA. 2016. Climate change indicators in the United States, 2016. Fourth Edition. EPA 430-R-16-004.

EPA. 2015. Climate Change in the United States: Benefits of Global Action. United States Environmental Protection Agency, Office of Atmospheric Programs, EPA 430-R-15-001.

Fears D. 2015. As salmon vanish in the dry Pacific Northwest, so does Native heritage. Washington Post.

Fisichelli et al (2015) Fisichelli NA, Schuurman GW, Monahan WB, Ziesler PS (2015) Protected Area Tourism in a Changing Climate: Will Visitation at US National Parks Warm Up or Overheat? PLoS ONE 10(6): e0128226. doi:10.1371/journal.pone.0128226

Fosu BO, Simon Wang S-Y, Yoon J-H. 2016. The 2014/15 Snowpack Drought in Washington State and its Climate Forcing. Bulletin of the American Meteorological Society 97(12): S19–S24. DOI: 10.1175/BAMS-D-16-0154.1.

Fritze JG, Blashki GA, Burke S, Wiseman J. 2008. Hope, despair and transformation: Climate change and the promotion of mental health and wellbeing. International Journal of Mental Health Systems.2:13.

Gamble, J.L., J. Balbus, M. Berger, K. Bouye, V. Campbell, K. Chief, K. Conlon, A. Crimmins, B. Flanagan, C. Gonzalez-Maddux, E. Hallisey, S. Hutchins, L. Jantarasami, S. Khoury, M. Kiefer, J. Kolling, K. Lynn, A. Manangan, M. McDonald, R. Morello-Frosch, M.H. Redsteer, P. Sheffield, K. Thigpen Tart, J. Watson, K.P. Whyte, and A.F. Wolkin, 2016: Ch. 9: Populations of Concern. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. U.S. Global Change Research Program, Washington, DC, 247–286. http://dx.doi.org/10.7930/J0Q81B0T

Garibaldi, A. and N. Turner. 2004. Cultural keystone species: implications for ecological conservation and restoration. Ecology and Society 9(3): 1. [online] URL: http://www.ecologyandsociety.org/vol9/iss3/art1/

Gentemann CL, Fewings MR, García-Reyes M. 2017. Satellite sea surface temperatures along the West Coast of the United States during the 2014–2016 northeast Pacific marine heat wave. Geophysical Research Letters 44(1): 2016GL071039. DOI: 10.1002/2016GL071039.

Goddard JHR, Treneman N, Pence WE, Mason DE, Dobry PM, Green B, Hoover C. 2016. Nudibranch Range Shifts Associated with the 2014 Warm Anomaly in the Northeast Pacific. Bulletin, Southern California Academy of Sciences 115(1): 15–40. DOI: 10.3160/soca-115-01-15-40.1.

Goode, J. R., J. M. Buffington, D. Tonina, D. J. Isaak, R. F. Thurow, S. Wenger, D. Nagel, C. Luce, DoertheTetzlaff, and C. Soulsby (2013), Potential effects of climate change on streambed scour and risks to salmonid survival in snow-dominated mountain basins, Hydrologic Processes, 27(5), 750–765, 10.1002/hyp.9728.

Gómez-Ocampo E, Gaxiola-Castro G, Durazo R, Beier E. (2017). Effects of the 2013-2016 warm anomalies on the California Current phytoplankton. Deep Sea Research Part II: Topical Studies in Oceanography. DOI: 10.1016/j.dsr2.2017.01.005.

Green M, Hamberg A, Main E, Early‐Alberts J, Dubuisson N, Douglas JP. 2013. Climate Smart Communities Scenarios Health Impact Assessment. Oregon Health Authority. April 2013: Portland, OR.

Haggerty B, York E, Early-Alberts J, Cude C. 2014. Oregon Climate and Health Profile Report. Oregon Health Authority: Portland, OR.

Haggerty B, 2015. Oregon Climate and Health Vulnerability Assessment. Oregon Health Authority: Portland, OR.

Haigh, R., Ianson, D., Holt, C. A., Neate, H. E., & Edwards, A. M. (2015). Effects of ocean acidification on temperate coastal marine ecosystems and fisheries in the northeast pacific. PLoS ONE, 10(2), 1–46. http://doi.org/10.1371/journal.pone.0117533

Haines, A., McMichael, A. J., Smith, K. R., Roberts, I., Woodcock, J., Markandya, A., ... & Bruce, N. (2010). Public health benefits of strategies to reduce greenhouse-gas emissions: overview and implications for policy makers. The Lancet, 374(9707), 2104-2114.

Hall et al. 2016. Columbia River Basin: Long-term water supply and demand forecast. DOE Pub No. 16-12-001.

Halofsky, J. E., and D. L. Peterson (2016), Climate change vulnerabilities and adaptation options for forest vegetation management in the northwestern USA, Atmosphere, 7(3), 46.

Halofsky, J. E., D. L. Peterson, K. L. Metlen, M. G. Myer, and V. A. Sample (2016), Developing and Implementing Climate Change Adaptation Options in Forest Ecosystems: A Case Study in Southwestern Oregon, USA, Forests, 7(11), 268.

Halofsky, J. E., D. L. Peterson, and H. R. Prendeville (2017), Assessing vulnerabilities and adapting to climate change in northwestern US forests, Climatic Change, 1-14.

Hatten, J. R., Waste, S. M., & Maule, A. G. (2014). Assessing climate-change risks to cultural and natural resources in the Yakima River Basin, Washington, USA. Climatic Change, 124(1–2), 363–370. http://doi.org/10.1007/s10584-014-1126-z

Hamlet, A.F. et al., 2013. An overview of the Columbia Basin Climate Change Scenarios Project: Approach, methods, and summary of key results. Atmosphere Ocean, 51(4), 392-415, doi: 10.1080/07055900.2013.819555

Hamlet, A. F. (2011). Assessing water resources adaptive capacity to climate change impacts in the Pacific Northwest Region of North America. Hydrology and Earth System Sciences, 15(5), 1427–1443. http://doi.org/10.5194/hess-15-1427-2011

Hamlet, A. F., and D. P. Lettenmaier (2007), Effects of 20th century warming and climate variability on flood risk in the western U.S., Water Resour. Res., 43(6), doi:10.1029/2006WR005099.

Hanson et al. 2014. Long-term impacts of major water storage facilities on agriculture and the natural environment: Evidence from Idaho. Ecological Economics 100:106-118.

HCCC. 2013. Hood Canal Integrated Watershed Plan: Goals and Guidance. https://hcccwagov.app.box.com/s/6a4qtpjpx2hwkd8xlqb7 (accessed May 21, 2017).

Hernandez, T., Gabbard, S., Carroll, D., 2016. A Demographic and Employment Profile of United States Farm workers. Findings from the National Agricultural Workers Survey (NAWS 2013-2014. U.S. Department of Labor. Employment and Training Administration. Office of Policy Development and Research. Research Report No. 12. https://www.doleta.gov/agworker/pdf/NAWS_Research_Report_12_Final_508_Compliant.pdf

Hessburg, P.F., Churchill, D.J., Larson, A.J. et al. Landscape Ecol (2015) 30: 1805. doi:10.1007/s10980-015-0218-0

Heindel JJ, Balbus J, Birnbaum L, Brune-Drisse MN, Grandjean P, Gray K, Landrigan PJ, Sly PD, Suk W, Slechta CD, Thompson C, Hanson M. (2016). Developmental Origins of Health and Disease: Integrating Environmental Influences. Endocrinology 17-22 doi: 10.1210/en.2015-1394

Hines JZ, Jagger MA, Jeanne TL, West N, Winquist A, Robinson B, Leman R, Hedberg K, 2017. Heavy Precipitation as a Risk Factor for Shigellosis among Homeless Persons during an Outbreak - Oregon 2015-2016. Centers for disease Control and Prevention, Oregon Health Authority, Multnomah County Health Department, Portland, OR.

Honea JM, McClure MM, Jorgensen JC, Scheuerell MD (2016) Assessing freshwater life-stage vulnerability of an endangered Chinook salmon population to climate change influences on stream habitat. Clim Res 71:127-137. https://doi.org/10.3354/cr01434

Houston, L., Capalbo, S., Seavert, C. et al. Climatic Change (2017). doi:10.1007/s10584-017-1951-y

Hu Z-Z, Kumar A, Jha B, Zhu J, Huang B. 2016. Persistence and Predictions of the Remarkable Warm Anomaly in the Northeastern Pacific Ocean during 2014–16. Journal of Climate 30(2): 689–702. DOI: 10.1175/JCLI-D-16-0348.1.

Idaho Foodbank (2016). Idaho Hunger Statistics. Retrieved at: https://idahofoodbank.org/media/for-the-media/food-insecurity-in-idaho/

Idaho State Department of Agriculture: Idaho Agriculture Facts 2015. www.agri.idaho.gov.

Inkley, D., M. Price, P. Glick, T. Losoff, and B. Stein (2013), Nowhere to Run: Big Game Wildlife in a Warming World, 32 pp., National Wildlife Federation.

Insley, M., and M. Lei (2007), Hedges and trees: incorporating fire risk into optimal decisions in forestry using a no-arbitrage approach, Journal of Agricultural and Resource Economics, 492-514.

Intergovernmental Panel on Climate Change (IPCC) 2012: Summary for Policymakers. In: Managing the Risks of Extreme Events and Disasters to Advance Climate Change Adaptation [Field, C.B., V. Barros, T.F. Stocker, D. Qin, D.J. Dokken, K.L. Ebi, M.D. Mastrandrea, K.J. Mach, G.-K. Plattner, S.K. Allen, M. Tignor, and P.M. Midgley (eds.)]. A Special Report of Working Groups I and II of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, UK, and New York, NY, USA, pp. 3-21. https://www.ipcc.ch/pdf/special-reports/srex/SREX_Full_Report.pdf

IOM (Institute of Medicine). 2007. Hospital based emergency care: At the breaking point. Washington, DC: The National Academies Press.

IOM (Institute of Medicine). 2010. Medical surge capacity: Workshop summary. Washington, DC: The National Academies Press.

Isaak, D. J., M. K. Young, C. H. Luce, S. W. Hostetler, S. J. Wenger, E. E. Peterson, J. M. Ver Hoef, M. C. Groce, D. L. Horan, and D. E. Nagel (2016), Slow climate velocities of mountain streams portend their role as refugia for cold-water biodiversity, Proceedings of the National Academy of Sciences, 113(16), 4374-4379, doi:10.1073/pnas.1522429113.

Isaak, D.J., S. Wollrab, D. Horan, and G. Chandler, (2012). Climate Change Effects on Stream and River Temperatures Across the Northwest U.S. from 1980-2009 and Implications for Salmonid Fishes. Climatic Change, 113(2), p. 499-524.

Isaak, D. J., C. H. Luce, B. E. Rieman, D. Nagel, E. Peterson, D. Horan, S. Parkes, and G. Chandler (2010), Effects of climate change and wildfire on stream temperatures and salmonid thermal habitat in a mountain river network, Ecological Applications, 20(5), 1350-1371, doi: 10.1890/09-0822.1.

Isaksen, T. B., Yost, M. G., Hom, E. K., Ren, Y., Lyons, H., & Fenske, R. A. (2015). Increased hospital admissions associated with extreme-heat exposure in King County, Washington, 1990-2010. Reviews on Environmental Health, 30(1), 51–64. http://doi.org/10.1515/reveh-2014-0050

Isaksen TB, Fenske RA, Hom EK, Ren Y, Lyons H, Yost MG. 2016. Increased mortality associated with extreme-heat exposure in King County, Washington, 1980–2010. International Journal of Biometeorology 60(1): 85–98. DOI: 10.1007/s00484-015-1007-9.

Izaurralde, Roberto C., et al. 2011. "Climate impacts on agriculture: implications for forage and rangeland production." Agronomy Journal 103.2 (2011): 371-381.

Jacox MG, Hazen EL, Zaba KD, Rudnick DL, Edwards CA, Moore AM, Bograd SJ. 2016. Impacts of the 2015-2016 El Nino on the California Current System: Early assessment and comparison to past events. Geophysical Research Letters 43(13): 7072–7080. DOI: 10.1002/2016GL069716.

Jackson, J. E., Yost, M. G., Karr, C., Fitzpatrick, C., Lamb, B. K., Chung, S. H., … Fenske, R. A. (2010). Public health impacts of climate change in Washington State: Projected mortality risks due to heat events and air pollution. Climatic Change, 102(1–2), 159–186. http://doi.org/10.1007/s10584-010-9852-3

Jaeger, W.J., A. Amos, D. Bigelow, H. Chang, D. Conklin, R. Haggerty, C. Langpap, K. Moore, P. Mote, A. Nolin, A.J. Plantinga, C. Schwartz, D. Tullos, and D. Turner, 201X: Climate change and water scarcity: surprises and insight fromn a regional system model. In review.

Jaffe DA, Zhang L. 2017. Meteorological anomalies lead to elevated O3 in the western U.S. in June 2015. Geophysical Research Letters 44(4): 2016GL072010. DOI: 10.1002/2016GL072010.

Jesdale, B. M. . M.-F., Rachel; Cushing, Lara, 2013. The racial/ethnic distribution of heat risk–related land cover in relation to residential segregation. Environmental Health Perspectives, 121, 811–817, doi:10.1289/ehp.1205919.

Jones, Gregory V. 2004. "Climate change in the western United States grape growing regions." VII International Symposium on Grapevine Physiology and Biotechnology 689.

Jones, R., Travers, C., Rodgers, C., Lazar, B., English, E., Lipton, J., … Martinich, J. (2013). Climate change impacts on freshwater recreational fishing in the United States. Mitigation and Adaptation Strategies for Global Change, 18(6), 731–758. http://doi.org/10.1007/s11027-012-9385-3

Joyce, LA et al. 2013. Climate Change and North American Rangelands: Assessment of Mitigation and Adaptation Strategies. Rangeland Ecology and Management 66:512-528.

Kais, S.M., Islam, M.S. 2016. Community capitals as community resilience to climate change: conceptual connections. International Journal of Environmental Research and Public Health. doi:10.3390/ijerph13121211

Kam J, Knutson TR, Zeng F, Wittenberg AT. 2016. Multimodel Assessment of Anthropogenic Influence on Record Global and Regional Warmth during 2015 [in “Explaining Extremes of 2015 from a Climate Perspective”]. Bulletin of the American Meteorological Society 97.

Karimi, T., et al. "Climate change and dryland wheat systems in the US Pacific Northwest." Agricultural Systems (2017).

Kaur et al. In review. Future bioclimatic drivers increase uncertainty of dryland agro-ecological classes of the IPNW, USA. Frontiers in Ecology and Evolution: Agroecology and Land Use Systems.

Keay JA, Peek JM (1980) Relationships between fires and winter habitat of deer in Idaho. Journal Wildlife Management 44: 372–380.

Kenney, B. 2012. Tribes as Managers of Federal Natural Resources. Natural Resources and Environment 27(1):1-4.

Kennedy, N.P., G.G. Yorgey, C.S. Frear, and C.E. Kruger. 2015. On-farm co-digestion of dairy manure with high energy organics. Washington State University Extension Publication FS172E, Pullman, WA.

Kennedy J, Morice C, Parker D, Kendon M. 2016. Global and regional climate in 2015. Weather 71(8): 185–192. DOI: 10.1002/wea.2760.

Klinsky, et al. 2016. Why equity is fundamental in climate change policy research, Global Environmental Change, http://dx.doi.org/10.1016/j.gloenvcha.2016.08.002

Kolb, T. E., C. J. Fettig, M. P. Ayres, B. J. Bentz, J. A. Hicke, R. Mathiasen, J. E. Stewart, and A. S. Weed (2016), Observed and anticipated impacts of drought on forest insects and diseases in the United States, Forest Ecology and Management, 380, 321-334, http://dx.doi.org/10.1016/j.foreco.2016.04.051.

Kormos, P., C. Luce, S. J. Wenger, and W. R. Berghuijs (2016), Trends and Sensitivities of Low Streamflow Extremes to discharge Timing and Magnitude in Pacific Northwest Mountain Streams, Water Resour. Res, 52(7), 4990-5007, 10.1002/2015WR018125.

Krakoff, S. 2012. Tribes, Land, and the Environment. Routledge, 248p.

Kronk-Warner, E. 2015. Everything Old Is New Again: Enforcing tribal Treaty Provisions to Protect Climate Change Threatened Resources. University of Kansas School of Law Working Paper. 41p.

Kruk, M.C., Marra, J.J., Merrifield, M., Ruggiero, P. Atkinson, D., Levinson, D., and Lander, M., 2013. Pacific Storms Climatology Products (PSCP): Understanding Extreme Events, Bulletin of the American Meteorological Society. doi: 10.1175/BAMS-D-11-00075.1

Kruger, C., E. Allen, J. Abatzoglou, K. Rajagopalan, and E. Kirby. 2017. Climate Considerations (ch. 1). In Yorgey, G. and C. Kruger, eds. Advances in Dryland Farming in the Inland Pacific Northwest, Washington State University Extension Publication EM108, Pullman, WA.

Lade S.J., Haider L.J., Engström G. and Schlüter M. 2017. Resilience offers escape from trapped thinking on poverty alleviation. Sci Adv 2017. (May 3, 2017) doi: 10.1126/sciadv.1603043

Lal, R. 2014. "Societal Value of Soil Carbon." Journal of Soil and Water Conservation 69(6):186A-92A.

Lane D, Jones R, Mills D, Wobus C, Ready RC, Buddemeier RW, English E, Martinich J, Shouse K, Hosterman H. (2015). Climate change impacts on freshwater fish, coral reefs, and related ecosystem services in the United States. Climatic Change 131(1): 143–157. DOI: 0.1007/s10584-014-1107-2.

Latif, Q. S., J. S. Sanderlin, V. A. Saab, W. M. Block, and J. G. Dudley. 2016. Avian relationships with wildfire at two dry forest locations with different historical fire regimes. Ecosphere 7:e01346. 10.1002/ecs2.1346

Latta, G., H. Temesgen, and T. M. Barrett (2009), Mapping and imputing potential productivity of Pacific Northwest forests using climate variables, Canadian Journal of Forest Research, 39(6), 1197-1207.

Liang, Xin-Zhong, et al. "Determining climate effects on US total agricultural productivity." Proceedings of the National Academy of Sciences 114.12 (2017): E2285-E2292.

Littell, J.S., Elsner, M.M., Whitely Binder, L.C., Snover, A.K. 2009. The Washington Climate Change Impacts Assessment: Evaluating Washington’s Future in a Changing Climate – Executive Summary. In The Washington Climate Change Impacts Assessment: Evaluating Washington’s Future in a Changing Climate, Climate Impacts Group, University of Washington, Seattle, Washington.

Leif E. Anderson and Mark L. Plummer, "Recreational Demand for Shellfish Harvesting Under Environmental Closures," Marine Resource Economics 32, no. 1 (January 2017): 43-57.

Littell, J. S., D. L. Peterson, K. L. Riley, Y. Liu, and C. H. Luce (2016), A review of the relationships between drought and forest fire in the United States, Global Change Biology, 22(7), 2353–2369, 10.1111/gcb.13275.

Liu JC, Mickley LJ, Sulprizio MP, Dominici F, Yue X, Ebisu K, Anderson GB, Khan RFA, Bravo MA, Bell ML. 2016. Particulate air pollution from wildfires in the Western US under climate change. Climatic Change 138(3–4): 655–666. DOI: 10.1007/s10584-016-1762-6.

Loomis, J., & Crespi, J. (1999). Estimated effects of climate change on selected outdoor recreation activities in the United States. In R. Mendelsohn & J. Neumann (Eds.), The Impact of Climate Change on the United States Economy (pp. 289-314). Cambridge: Cambridge University Press. doi:10.1017/CBO9780511573149.012

Luce, C. H., J. M. Vose, N. Pederson, J. Campbell, C. Millar, P. Kormos, and R. Woods (2016), Contributing Factors for Drought in United States Forest Ecosystems under Projected Future Climates and their Uncertainty, Forest Ecology and Management, 380, 299-308, 10.1016/j.foreco.2016.05.020.

Luce, C., B. Staab, M. Kramer, S. Wenger, D. Isaak, and C. McConnell (2014b), Sensitivity of Summer Stream Temperatures to Climate Variability in the Pacific Northwest, Water Resour. Res, 50(4), 3428-3443, DOI: 10.1002/2013WR014329.

Luce, C. H., J. T. Abatzoglou, and Z. A. Holden (2013), The Missing Mountain Water: Slower Westerlies Decrease Orographic Enhancement in the Pacific Northwest USA, Science, 342(6164), 1360-1364, DOI: 10.1126/science.1242335.

Lynn, K., Daigle, J., Hoffman, J., Lake, F., Michelle, N., Ranco, D., Viles, C., Voggesser, G. Williams, P. (2013). The impacts of climate change on tribal traditional foods. Climatic Change, 120, 545–556. http://doi.org/10.1007/s10584-013-0736-1

Maizlish, N., Woodcock, J., Co, S., Ostro, B., Fanai, A., & Fairley, D. (2013). Health cobenefits and transportation-related reductions in greenhouse gas emissions in the San Francisco Bay area. American Journal of Public Health, 103(4), 703-709.

Mantua, N., Tohver, I., & Hamlet, A. (2010). Climate change impacts on streamflow extremes and summertime stream temperature and their possible consequences for freshwater salmon habitat in Washington State. Climatic Change, 102(1–2), 187–223. http://doi.org/10.1007/s10584-010-9845-2

Margulis SA, Cortes G, Girotto M, Huning LS, Li D, Durand M. 2016. Characterizing the extreme 2015 snowpack deficit in the Sierra Nevada (USA) and the implications for drought recovery. Geophysical Research Letters 43(12): 6341–6349. DOI: 10.1002/2016GL068520.

Margles Weis SW, Agostini VN, Roth LM, Gilmer B, Schill SR, Knowles JE, Blyther R, 2016. Assessing Vulnerability: An integrated approach for mapping adaptive capacity, sensitivity, and exposure. Climatic Change, Vol. 136, Issue 3, pp 615-629. DOI: 10.1007/s10584-016-1642-0

Marris E. 2015. The Klamath Tribes Tribal News and Events. In the Dry West, Waiting for Congress.

Mauger, Guillaume, Yoram Bauman, Tamilee Nennich, and Eric Salathé. 2015a. "Impacts of climate change on milk production in the United States." The Professional Geographer 67, no. 1: 121-131.

Mauger, G.S., J.H.Casola, H.A.Morgan, R.L.Strauch, B.Jones, B.Curry, T.M.BuschIsaksen, L. Whitely Binder, M.B. Krosby, and A.K. Snover, 2015b. State of Knowledge: Climate Change in Puget Sound. Report prepared for the Puget Sound Partnership and the National Oceanic and Atmospheric Administration. Climate Impacts Group, University of Washington, Seattle. doi:10.7915/CIG93777D
http://cses.washington.edu/picea/mauger/ps-sok/PS-SoK_2015.pdf

McCabe RM, Hickey BM, Kudela RM, Lefebvre KA, Adams NG, Bill BD, Gulland FMD, Thomson RE, Cochlan WP, Trainer VL. 2016. An unprecedented coastwide toxic algal bloom linked to anomalous ocean conditions. Geophysical Research Letters 43(19): 2016GL070023. DOI: 10.1002/2016GL070023.

McCarl et al. 2016. The challenge of climate change adaptation for agriculture: An economically oriented review. J. Agricultural and Applied Economics 48(4): 321-344.

McKelvey, K. S., J. P. Copeland, M. K. Schwartz, J. S. Littell, K. B. Aubry, J. R. Squires, S. A. Parks, M. M. Elsner, and G. S. Mauger (2011), Climate change predicted to shift wolverine distributions, connectivity, and dispersal corridors, Ecological Applications, 21(8), 2882-2897.

McKenzie and Littell (2017). Climate Change and the eco-hydrology of fire: Will area burned increase in a warming western USA? Ecological Applications 21(1):26-36.

McLain K, Hancock J, Drennan M. 2017. 2015 Drought and Agriculture. Washington State Department of Agriculture: Olympia, WA.

McNeeley, S. M. (2017). Sustainable Climate Change Adaptation in Indian Country. Weather, Climate, and Society, 393–404. http://doi.org/10.1175/WCAS-D-16-0121.1

McOliver, C., Camper, A., Doyle, J., Eggers, M., Ford, T., Lila, M., Berner, J., Campbell, L., and Donatuto, J., 2015. Community-Based Research as a Mechanism to Reduce Environmental Health Disparities in American Indian and Alaska Native communities. International Journal of Environmental Research and Public Health, 2015 (12). ISSN 1660-4601 doi: 10.3390/ijerph120404076.

Melillo, Jerry M., Terese (T.C.) Richmond, and Gary W. Yohe, Eds., 2014: Climate Change Impacts in the United States: The Third National Climate Assessment, Chapter 12 Indigenous Peoples. U.S. Global Change Research Program, 841 pp. doi:10.7930/J0Z31WJ2.

Messervy, M. (2016) Insurer Climate Risk Disclosure Survey Report & Scorecard: 2016 FINDINGS & RECOMMENDATIONS. Report from Ceres. https://www.ceres.org/resources/reports/2016-insurer-climate-risk-disclosure-survey/view

Milstein M. 2015. NOAA Northwest Fisheries Science Center. NOAA Fisheries mobilizes to gauge unprecedented West Coast toxic algal bloom.

Montag, J. M., Swan, K., Jenni, K., Nieman, T., Hatten, J., Mesa, M., Graves, D., Voss, F., Mastin, M., Hardiman, J., Maule, A. (2014). Climate change and Yakama Nation tribal well-being. Climatic Change, 124(1–2), 385–398. http://doi.org/10.1007/s10584-013-1001-3

Moore SK, Johnstone JA, Banas NS, Salathé Jr. EP. 2015. Present-day and future climate pathways affecting Alexandrium blooms in Puget Sound, WA, USA. Harmful Algae 48: 1–11. DOI: 10.1016/j.hal.2015.06.008.

Morello-Frosch, R., Zuk, M., Jerrett, M., Shamasunder, B., Kyle, A., 2011. Understanding the Cumulative Impacts of Inequalities in Environmental Health: Implications for Policy. Health Affairs. May 2011. Vol 30 No 5 879-887. doi: 10.1377/hlthaff.2011.0153 http://content.healthaffairs.org/content/30/5/879.full

Morello-Frosch, R., Pastor, M., Sadd, J., Shankoff, S., 2009. The Climate Gap: Inequalities in How Climate Change Hurts Americans & How to Close the Gap. College of Natural Resources, University of California, Berkeley. Department of Environmental Science, Policy and Management. https://dornsife.usc.edu/pere/climategap/

Morrow et al. 2016. Climate change predicted to negatively influence surface soil organic matter of dryland cropping systems in the IPNW, USA. Frontiers in Ecology and Evolution: Agroecology and Land Use Systems 5:10.

Morton, Lois Wright, Walt Mahaffee and Mark Gleason. 2017a. Climate, Weather and Wine Grapes. Sociology Technical Report 1043. Department of Sociology, Iowa State University, Ames, Iowa. 18 pp.

Morton, Lois Wright, David Gent and Mark Gleason. 2017b. Climate, Weather and Hops. Sociology Technical Report 1045. Department of Sociology, Iowa State University, Ames, Iowa. 24 pp.

Mote, P.W., et al., 2017.

Mote, P.W., D.Rupp, S. Li, F. Otto, D. Sharp, F. Otto, P. Uhe, M. Xiao, D. Lettenmaier, H. Cullen, and M.R. Allen, 2016: Perspectives on the causes of exceptionally low 2015 snowpack in the western US. Geophys. Research Letters, 43, doi:10.1002/ 2016GL069965.

Mote, P., A. K. Snover, S. Capalbo, S. D. Eigenbrode, P. Glick, J. Littell, R. Raymondi, and S. Reeder, 2014: Ch. 21: Northwest. Climate Change Impacts in the United States: The Third National Climate Assessment, J. M. Melillo, Terese (T.C.) Richmond, and G. W. Yohe, Eds., U.S. Global Change Research Program, 487-513. doi:10.7930/J04Q7RWX.

Mote, P.W., and 18 co-authors, 1999: Impacts of climate variability and change: Pacific Northwest. A report of the JISAO/SMA Climate Impacts Group. 110 pp

Multnomah County and City of Portland. 2014. Climate Change Preparation Strategy - Risk and Vulnerabilities Assessment. https://multco.us/file/36550/download

Myers,S.S., Smith M.R. , Guth S., Golden CD. , Vaitla B., Mueller N.D. , Dangour AD. ,Huybers P. Climate Change and Global Food Systems: Potential Impacts on Food Security and Undernutrition. Annual Review of Public Health. 2017. 10.1146/annurev-publhealth-031816-044356

Nehlsen, W., J. Williams, and J. Lichatowich. 1991. Pacific Salmon at the Crossroads: Stocks at Risk from California, Oregon, Idaho, and Washington. Fisheries 16(2):4-21. https://www.researchgate.net/publication/232832878_Pacific_Salmon_at_the_Crossroads_Stocks_at_Risk_From_California_Oregon_Idaho_and_Washington

Nguyen T, Davis K, Counts N, Fritze D, 2016. The State of Mental Health in America. Mental Health America. http://www.mentalhealthamerica.net/issues/state-mental-health-america

Niebergs et al. In Press. Estimating climate change effects on grazing management and beef cattle production in the Pacific Northwest, Climatic Change.

Niemi, E.G. Buckley, M., Neculae, C., Reich, S., & Climate Leadership Initiative (2009) “An Overview of Potential Economic Costs to Washington of a Business-As-Usual Approach to Climate Change”, ECONorthwest, 47 pp.

NOAA Fisheries. 2015. “Farmers Reroute Water for Fish” NOAA Fisheries West Coast Region. Accessed September 19, 2017. http://www.westcoast.fisheries.noaa.gov/stories/2015/21_07212015_yakima_canal_flows.html.

NOAA National Centers for Environmental information, Climate at a Glance: U.S. Time Series, Average Temperature, published April 2017, retrieved on April 11, 2017 from http://www.ncdc.noaa.gov/cag/

NOAA National Centers for Environmental Information, State of the Climate: Global Climate Report for Annual 2016, published online January 2017, retrieved on April 11, 2017 from https://www.ncdc.noaa.gov/sotc/global/201613.

NOAA State Climate Summaries for OR, WA, and ID: https://statesummaries.ncics.org

Norgaard, K.M. (2005). The effects of altered diet on the health of the Karuk people. http://pages.uoregon.edu/norgaard/pdf/Effects-Altered-Diet-Karuk-Norgaard-2005.pdf.

Northwest Harvest (2015). Washington Hunger Facts. Retrieved at: http://www.northwestharvest.org/wa-hunger-facts

Northwest Power and Conservation Council (NWPCC). 2016. Seventh Power Plan: Energy efficiency, demand response, and existing natural gas plants key to ensuring a reliable, low-cost, clean power supply. https://www.nwcouncil.org/energy/powerplan/7/plan

Norton-Smith, Kathryn; Lynn, Kathy; Chief, Karletta; Cozzetto, Karen; Donatuto, Jamie; Hiza Redsteer, Margaret; Kruger, Linda E.; Maldonado, Julie; Viles, Carson; Whyte, Kyle P. 2016. Climate change and indigenous peoples: a synthesis of current impacts and experiences. Gen. Tech. Rep. PNW-GTR-944. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 136 p. https://www.fs.fed.us/pnw/pubs/pnw_gtr944.pdf

Ocean Park Food Bank, 2017 http://www.chinookobserver.com/co/local-news/20170531/domoic-acid-hurt-jobs-along-with-clams

O’Neal, K. (2002). Effects of Global Warming on Trout and Salmon in U.S. Streams. Defenders of Wildlife and Natural Resources Defense Council. May.

Olen, Beau, JunJie Wu, and Christian Langpap. 2016. "Irrigation decisions for major west coast crops: Water scarcity and climatic determinants." American Journal of Agricultural Economics 98.1 (2016): 254-275.

Oregon Department of Agriculture (ODA): Oregon Agriculture Facts and Figures, August 2016. www.oregon.gov/ODA

Oregon Department of Transportation and Oregon Health Authority (ODOT and OHA) 2016. How Tillamook Weathered the Storm: A Case Study on Creating Climate Resilience on Oregon’s North Coast. Oregon Department of Transportation and Oregon Health Authority. http://www.oregon.gov/oha/ph/HealthyEnvironments/climatechange/Documents/Tillamook-Case-Study-DRAFT-12-13-16b.pdf

Oregon Food Bank (2016). Hunger factors 2015: Hunger and Poverty in Oregon and Clark County, WA. Retrieved at: http://www.oregonfoodbank.org/wp-content/uploads/2016/05/Hunger-Factors-FullRpt-v8-2.pdf

Oregon Health Authority (OHA) 2015. Social Vulnerability - Populations Sensitivity - Composite Vulnerability Index. July 2015

Oregon Health Authority, 2016. State of Oregon Public Health Modernization Assessment Report. Portland, OR.

Oregon Public Health Division, Acute & Communicable Disease Prevention, 2015. Oregon ESSENCE Hazard Report. Portland, OR.

Oregon Public Health Division, Oregon Health Authority 2013. CD Summary: Racial and Ethnic Health Disparities in Oregon.

Ou, M., Hamilton, T. J., Eom, J., Lyall, E. M., Gallup, J., Jiang, A., … Brauner, C. J. (2015). Responses of pink salmon to CO2-induced aquatic acidification. Nature Climate Change, 5(June), 1–24. http://doi.org/10.1038/nclimate2694

Outdoor Recreation Association (2012) “The Outdoor Recreation Economy” https://outdoorindustry.org/research-tools/outdoor-recreation-economy/

Pacifici, M and Piero Visconti, Stuart H. M. Butchart, James E. M. Watson, Francesca M. Cassola & Carlo Rondinini. 2017. Species’ traits influenced their response to recent climate change. Nature Climate Change 7:205-208.

Paek H, Yu J-Y, Qian C. 2017. Why were the 2015/2016 and 1997/1998 extreme El Ninos different? Geophysical Research Letters 44(4): 1848–1856. DOI: 10.1002/2016GL071515.

Paerl, H. W. and Huisman, J. (2009), Climate change: a catalyst for global expansion of harmful cyanobacterial blooms. Environmental Microbiology Reports, 1: 27–37. doi:10.1111/j.1758-2229.2008.00004.

Papiez, C. (2009). Climate Change Implications for the Quileute and Hoh Tribes of Washington: A Multidisciplinary Approach to Assessing Climatic Disruptions to Coastal Indigenous Communities.

Parker and Abatzoglou 2016. Projected changes to cold hardiness zones and suitable overwinter ranges of perennial crops over the United States. Environmental Research Letters 11: 034001.

Peden A, (2016). Oregon CCO Housing Supports: Survey Report. Oregon Health Authority: http://www.oregon.gov/oha/HPA/HP/docs/OHA%208440%20CCO-Housing-Survey-Report.pdf

Perera FP, 2017. Multiple Threats to Child Health from Fossil Fuel Combustion: Impacts of Air Pollution and Climate Change. Environmental Health Perspectives. Feb; 125(2): 141-148. doi: 10.1289/EHP299

Peterson, D. L., J. M. Vose, and T. Patel-Weynand (2013), Climate change and United States forests, Springer.

Peterson, B., Rauh, V., Bansal, R., Hao, X., Toth, Z., Nati, G., . . . Perera, F. (2015). Effects of prenatal exposure to air pollutants (polycyclic aromatic hydrocarbons) on the development of brain white matter, cognition, and behavior in later childhood. JAMA Psychiatry, 531-40. doi:10.1001/jamapsychiatry.2015.57

Peterson, D. L., C. I. Millar, L. A. Joyce, M. J. Furniss, J. E. Halofsky, R. P. Neilson, and T. L. Morelli (2011), Responding to climate change in national forests: a guidebook for developing adaptation options. Gen. Tech. Rep. PNW-GTR-855., 109 pp., U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station, Portland, OR.

Peterson W, Bond N, Robert M. 2016. The Blob (Part Three): Going, going, gone? PICES Press 24(1): 46–48.

Poe, M. R., Donatuto, J., & Satterfield, T. (2016). “Sense of Place”: Human Wellbeing Considerations for Ecological Restoration in Puget Sound. Coastal Management, 44(5), 409–426. http://doi.org/10.1080/08920753.2016.1208037

Poe, M. R., Levin, P. S., Tolimieri, N., & Norman, K. (2015). Subsistence fishing in a 21st century capitalist society: From commodity to gift. Ecological Economics, 116, 241–250. http://doi.org/10.1016/j.ecolecon.2015.05.003

Poe, M. R., Norman, K. C., & Levin, P. S. (2014). Cultural dimensions of socioecological systems: Key connections and guiding principles for conservation in coastal environments. Conservation Letters, 7(3), 166–175. http://doi.org/10.1111/conl.12068

Poeplau, Christopher and Axel Don. "Carbon sequestration in agricultural soils via cultivation of cover crops–A meta-analysis." Agriculture, Ecosystems & Environment 200 (2015): 33-41.

Polley et al. 2013. Climate Change and North American Rangelands: Trends, Projections, and Implications. Rangeland Ecol Manage 66:493-511.

Portland and Multnomah County. 2016. Climate Action Through Equity. The integration of equity in the Portland/Multnomah County 2015 Climate Action Plan. July 12, 2016. https://www.portlandoregon.gov/bps/article/583501

Portland and Multnomah County. 2015. Portland-Multnomah Climate Action Plan. June, 2015. https://www.portlandoregon.gov/bps/article/531984

Prochaska JD, Nolen AB, Kelley H, Sexton K, Linder SH, Sullivan J (2014). Social Determinants of Health in Environmental Justice Communities: Examining Cumulative Risk in Terms of Environmental Exposures and Social Determinants of Health. Human and Ecological Risk Assessment: Vol. 20, - Issue 4, 980-994. DOI: 10.1080/10807039.2013.805957

Quesada, J., Hart, L., Bourgois, P., 2011. Structural Vulnerability and Health: Latino Migrant Laborers in the United States. Med Anthropol. Jul: 30(4): 339-362. doi: 10.1080/01459740.2011.576725

Rajagopalan, K, K. Chinayakanahalli, C. Stockle, R. Nelson, C. Kruger, M. Brady, K. Malek, S. Dinesh, M.E. Barber, A. Hamlet, G. Yorgey, and J. Adam. In review. Impacts of near term regional climate change on agricultural production in the Columbia River Basin. Water Resources Research.

Ranco DJ, O’Neill CA, Donatuto J, Harper B. 2011. Environmental Justice, American Indians and the Cultural Dilemma: Developing Environmental Management for Tribal Health and Well-Being. Environmental Justice. December 2011, 4(3): 221-230 doi: 10.1089/env.2010.0036

Raymond, 2015. Seattle City Light Climate Change Vulnerability Assessment and Adaptation Plan. https://www.seattle.gov/light/enviro/docs/Seattle_City_Light_Climate_Change_Vulnerability_Assessment_and_Adaptation_Plan.pdf

Riedel, J. L., and coauthors. 2015. Glacier status and contribution to streamflow in the Olympic Mountains, Washington, USA. Journal of Glaciology 61(225):8-16

Rieman, B. E., D. J. Isaak, S. Adams, D. Horan, D. Nagel, C. Luce, and D. Myers (2007), Anticipated Climate Warming Effects on Bull Trout Habitats and Populations across the Interior Columbia River Basin, Trans. Am. Fish. Soc., 136(6), 1552-1565.

Ritokova, G., D.C. Shaw, G. Filip, A. Kanaskie, J. Browning, and D. Norlander (2016). Swiss Needle Cast in Western Oregon Douglas-Fir Plantations: 20‐Year Monitoring Results, Forests 7(8), 155; doi:10.3390/f7080155.

Roser-Renouf, C., Maibach, E. W., & Li, J. (2016). Adapting to the changing climate: An assessment of local health department preparations for climate change-related health threats, 2008-2012. PloS one, 11(3), e0151558.

Rudolph, Linda, et al. 2013. "Health in All Policies: improving health through intersectoral collaboration." National Academy of Sciences.

Ruggiero, Peter, Kratzmann, M.G., Himmelstoss, E.A., Reid, David, Allan, John, and Kaminsky, George, 2013, National assessment of shoreline change—Historical shoreline change along the Pacific Northwest coast: U.S. Geological Survey Open-File Report 2012–1007, 62 p., http://dx.doi.org/10.3133/ofr20121007. ISSN 0196-1497 (print); ISSN 2331-1258 (online)

Ruggiero., P., 2013. Is the intensifying wave climate of the U.S. Pacific Northwest increasing flooding and erosion risk faster than sea level rise?, Journal of Waterway, Port, Coastal, and Ocean Engineering. 10.1061/(ASCE)WW.1943-5460.0000172

Rupp, D.E., J.T. Abatzoglou, and P.W Mote, 2016: Projections of 21st century climate of the Columbia River Basin. Clim. Dyn., doi: 10.1007/s00382-016-3418-7.

Saab, Victoria A.; Latif, Quresh S.; Rowland, Mary M.; Johnson, Tracey N.; Chalfoun, Anna D.; Buskirk, Steven W.; Heyward, Joslin E.; Dresser, Matthew A. 2014. Ecological Consequences of Mountain Pine Beetle Outbreaks for Wildlife in Western North American Forests. Forest Science 60:539-559.

Safeeq, M., Mauger, G. S., Grant, G. E., Arismendi, I., Hamlet, A. F., & Lee, S.-Y. (2014). Comparing Large-Scale Hydrological Model Predictions with Observed Streamflow in the Pacific Northwest: Effects of Climate and Groundwater*. Journal of Hydrometeorology, 15(6), 2501–2521. http://doi.org/10.1175/JHM-D-13-0198.1

Sakuma KM, Field JC, Mantua NJ, Ralston S, Marinovic BB, Carrion CN. 2016. Anomalous Epipelagic Micronekton Assemblage Patterns in the Neritic Waters of the California Current in Spring 2015 During a Period of Extreme Ocean Conditions. California Cooperative Oceanic Fisheries Investigations Reports 57: 163–183.

Sanford, E. Water Temperature, Predation, and the Neglected Role of Physiological Rate Effects in Rocky Intertidal Communities. Integr Comp Biol 2002; 42 (4): 881-891. doi: 10.1093/icb/42.4.881

Saperstein, A. 2015. Climate Change, Migration, and the Puget Sound Region: What We Know and How We Could Learn More. Report prepared for the University of Washington Climate Impacts Group. The Daniel J. Evans School of Public Policy and Governance, University of Washington, Seattle.

Sarofim MC, Saha S, Hawkins MD, Mills DM, Hess J, Horton R, Kinney P, Schwartz J, St. Juliana A. 2016. Ch. 2: Temperature-Related Death and Illness. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. US Global Change Research Program: Washington, DC, 43–68.

Schaible and Aillery. 2017. Challenges for US Irrigated Agriculture in the Face of Emerging Demands and Climate Change. See ERS.

Schwartz JD, Lee M, Kinney PL, Yang S, Mills D, Sarofim MC, Jones R, Streeter R, Juliana AS, Peers J, Horton RM. 2015. Projections of temperature-attributable premature deaths in 209 U.S. cities using a cluster-based Poisson approach. Environmental Health 14: 85. DOI: 10.1186/s12940-015-0071-2.

Seattle City Light (SCL). 2015. Climate Change Vulnerability Assessment and Adaptation Plan. http://www.seattle.gov/light/enviro/docs/Seattle_City_Light_Climate_Change_Vulnerability_Assessment_and_Adaptation_Plan.pdf

Seattle City Light (SCL). 2016. Opportunities and Challenges. Utility of the Future. 2015 Annual Report, Financial Information. http://www.seattle.gov/light/pubs/annualrpt/2015/flipbook/index.html?page=14

Seattle, 2016a, Equity and Environment Agenda 2016. http://www.seattle.gov/Documents/Departments/OSE/SeattleEquityAgenda.pdf

Seattle, 2016b. Seattle 2035 Comprehensive Plan: Managing Growth to Become and Equitable and Sustainable City 2015-2035. November 2016. http://www.seattle.gov/dpd/cs/groups/pan/@pan/documents/web_informational/p2580242.pdf

Seattle, 2017 [forthcoming]. Preparing for Climate Change. Available June 2017. Office of Sustainability & Environment http://www.seattle.gov/Documents/Departments/OSE/ClimateDocs/SEAclimatePreparedness_Draft_Oct2016.pdf [need to update when final is released]

Sexton T, Perkins J, Rogers G, Kerr D, Engleman D, Wall D, Swedberg T, Pence M, Peterson J, Graw R, Murphy K, Strawn K. 2016. Narrative Timeline of the Pacific Northwest 2015 Fire Season. United States Forest Service Rocky Mountain Research Station: Fort Collins, CO, 281.

Shi, L., Chu, E., Anguelovski, I., Aylett, A., Debats, J., Goh, K., … VanDeveer, S. D. (2016). Roadmap towards justice in urban climate adaptation research. Nature Climate Change, 6(2), 131–137. http://doi.org/10.1038/nclimate2841

Sheridan, S. C. . K., Adam J.; Kalkstein, Laurence S., 2009: Trends in heat-related mortality in the United States, 1975–2004. Natural Hazards, 50, 145–160, doi:10.1007/s11069-008-9327-2.

Shonkoff, S. B., R. Morello-Frosch, M. Pastor, and J. Sadd, 2011. The climate gap: Environmental health and equity implications of climate change and mitigation policies in California—a review of the literature. Climatic Change, 109, 485–503, doi:10.1007/s10584-011-0310-7.

Scigliano, E (2012). Sweetening the Waters The Feasibility and Efficacy of Measures to Protect Washington’s Marine Resources from Ocean Acidification. For the Global Ocean Health Program, a joint project of the National Fisheries Conservation Center and the Sustainable Fisheries Partnership. http://www.ecy.wa.gov/water/marine/oa/2012report_app9.pdf

Siddika, N., Balogun, H., Amegah, A., & Jaakkola, J. (2016). Prenatal ambient air pollution exposure and the risk of stillbirth: systematic review and meta-analysis of the empirical evidence. Occup Environ Med., 573-81. doi:10.1136/oemed-2015-103086

Sims, C. (2011), Optimal timing of salvage harvest in response to a stochastic infestation, Natural Resource Modeling, 24(3), 383-408.

Smith, C.L. (1980). Attitudes about the value of steelhead and salmon angling. T Am. Fish Soc., 109, 272-281.

Snover, A.K, G.S. Mauger, L.C. Whitely Binder, M. Krosby, and I. Tohver. 2013. Climate Change Impacts and Adaptation in Washington State: Technical Summaries for Decision Makers. State of Knowledge Report prepared for the Washington State Department of Ecology. Climate Impacts Group, University of Washington, Seattle.

Sorte, B. and M. Rahe, and P. Lewin. 2016. Agriculture, Food, Forestry and Fishing in the NW U.S.: An Economic Overview. Oregon State University Extension Service Rural Studies Program and University of Idaho Extension Service with support from Northwest Farm Credit Services.

Sproles EA, Roth TR, Nolin AW. 2017. Future snow? A spatial-probabilistic assessment of the extraordinarily low snowpacks of 2014 and 2015 in the Oregon Cascades. The Cryosphere 11(1): 331–341. DOI: 10.5194/tc-11-331-2017.

Stanford, E. (2002). “Community Responses to Climate Change: Links Between Temperature and Keystone Predation in a Rocky Intertidal System.” In S.H. Schneider and T.L. Root (eds.). Wildlife Responses to Climate Change: North American Case Studies. Island Press: Washington D.C.

Stevenson J. 2016. Documenting the Drought. The Climate CIRCulator.

Stöckle, C.O., Higgins, S., Nelson, R. et al. Climatic Change (2017). doi:10.1007/s10584-017-1950-z

Stöckle, Claudio O., Roger L. Nelson, Stewart Higgins, Jay Brunner, Gary Grove, Rick Boydston, Matthew Whiting, and Chad Kruger. 2010. "Assessment of climate change impact on Eastern Washington agriculture." Climatic Change 102, no. 1 : 77-102.

Sun, X., Luo, X., Zhao, C., Chung Ng, R., Lim, C., Zhang, B., & Liu, T. (2015). The association between fine particulate matter exposure during pregnancy and preterm birth: a meta-analysis. BMC Pregnancy Childbirth. doi:10.1186/s12884-015-0738-2

Teague, W. R., et al. (2016) "The role of ruminants in reducing agriculture's carbon footprint in North America." Journal of Soil and Water Conservation 71.2 156-164.

Thornton, T., Deur, D., & Kitka, H. (2015). Cultivation of salmon and other marine resources on the Northwest Coast of North America. Human Ecology, 43(2), 189–199. http://doi.org/10.1007/s10745-015-9747-z

Travis B. Paveglio, Tony Prato, Catrin Edgeley, Darek Nalle. 2016. Evaluating the Characteristics of Social Vulnerability to Wildfire: Demographics, Perceptions, and Parcel Characteristics. Environmental Management. September 2016, Volume 58, Issue 3, pp 534–548

Trtanj J, Jantarasami L, Brunkard J, Collier T, Jacobs J, Lipp E, McLellan S, Moore S, Paerl H, Ravenscroft J, Sengco M, Thurston J. 2016. Ch. 6: Climate Impacts on Water-Related Illness. The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. U.S. Global Change Research Program: Washington, DC, 157–188.

Uejio, C. K., O. V. Wilhelmi, J. S. Golden, D. M. Mills, S. P. Gulino, and J. P. Samenow, 2011. Intra-urban societal vulnerability to extreme heat: The role of heat exposure and the built environment, socioeconomics, and neighborhood stability. Health & Place, 17, 498–507, doi:10.1016/j.healthplace.2010.12.005.

Urban Sustainability Directors Network (USDN) Innovation Fund. 2017. Guide to Equitable, Community-Driven Climate Preparedness Planning, Raimi + Associates and USDN. May 2017. https://www.usdn.org/public/page/18/Climate-Change-Preparedness#GuideClimate

U.S. Federal Government, 2014: U.S. Climate Resilience Toolkit. http://toolkit.climate.gov.

U.S. Climate Resilience Toolkit, 2017. Quinault Indian Nation Plans for Village Relcoation (accessed June 16, 2017) https://toolkit.climate.gov/case-studies/quinault-indian-nation-plans-village-relocation

U.S. Department of Energy (DOE). 2015a. Climate Change and the U.S. Energy Sector: Regional Vulnerabilities and Resilience Solutions. U.S. Department of Energy.

U.S. Department of Energy (DOE). 2015b. Quadrennial Energy Review Report: Energy Transmission, Storage, and Distribution Infrastructure. April 2015. https://energy.gov/sites/prod/files/2015/04/f22/QER-ALL%20FINAL_0.pdf

U.S. Department of Energy (DOE) 2013. Effects of Climate Change on Federal Hydropower - Report to Congress, US Department of Energy, Washington, DC.

U.S. Department of Health and Human Services. 2013. National Healthcare Disparities Report. Agency for Healthcare Research and Quality. Publication No. 14-0006. May 2014. https://archive.ahrq.gov/research/findings/nhqrdr/nhdr13/2013nhdr.pdf

U.S. Department of Housing and Urban Development. The 2016 Annual Homeless Assessment Report (AHAR) to Congress, Part 1 Point-in-Time Estimates of Homelessness. https://www.hudexchange.info/resources/documents/2016-AHAR-Part-1.pdf

U.S. Department of the Interior, U.S. Fish and Wildlife Service, and U.S. Department of Commerce, U.S. Census Bureau. 2011 National Survey of Fishing, Hunting, and Wildlife-Associated Recreation. https://www.census.gov/prod/2012pubs/fhw11-nat.pdf

USGCRP, 2016: The Impacts of Climate Change on Human Health in the United States: A Scientific Assessment. Crimmins, A., J. Balbus, J.L. Gamble, C.B. Beard, J.E. Bell, D. Dodgen, R.J. Eisen, N. Fann, M.D. Hawkins, S.C. Herring, L. Jantarasami, D.M. Mills, S. Saha, M.C. Sarofim, J. Trtanj, and L. Ziska, Eds. U.S. Global Change Research Program, Washington, DC, 312 pp. http://dx.doi.org/10.7930/J0R49NQX

Vano, Julie A., Michael J. Scott, Nathalie Voisin, Claudio O. Stöckle, Alan F. Hamlet, Kristian EB Mickelson, Marketa McGuire Elsner, and Dennis P. Lettenmaier. 2010. "Climate change impacts on water management and irrigated agriculture in the Yakima River Basin, Washington, USA." Climatic Change 102, no. 1 (2010): 287-317.

Venterea, Rodney T., et al. 2012. "Challenges and opportunities for mitigating nitrous oxide emissions from fertilized cropping systems." Frontiers in Ecology and the Environment 10.10 (2012): 562-570.

Vincent, G.K. and V.A. Velkoff. 2010. The Next Four Decades: The Older Population in the United States: 2010 to 2050 – Population Estimates and Projections [Internet]. Washington, DC: US Census Bureau. Current Population Reports: P25-1138

Vinyeta, K and K. Lynn. 2013. Exploring the Role of Traditional Ecological Knowledge in Climate Change Initiatives. Gen. Tech. Rep. PNW-GTR-879. Portland, OR: U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station. 43 p. https://www.fs.fed.us/pnw/pubs/pnw_gtr879.pdf

Vinyeta, Kirsten; Whyte, Kyle Powys; Lynn, K. (2015). Climate Change Through an Intersectional Lens: Gendered Vulnerability and Resilience in Indigenous Communities in the United States. Usda, (December), 88. Retrieved from https://www.researchgate.net/profile/Kyle_Whyte/publication/278675505_Climate_Change_Through_an_Intersectional_Lens_Gendered_Vulnerability_and_Resilience_in_Indigenous_Communities_in_the_United_States/links/5582cb8808ae1b14a0a27c43.pdf

Voggesser, G., Lynn, K., Daigle, J., Lake, F. K., & Ranco, D. (2013). Cultural impacts to tribes from climate change influences on forests. Climatic Change, 120, 615–626. http://doi.org/10.1007/s10584-013-0733-4

Vose et al. 2016 Effects of Drought on Forests and Rangelands in the United States: A Comprehensive Science Synthesis, U.S. Forest Service GTR WO-93b. See individual chapters

Vose, J. M., C. F. Miniat, C. H. Luce, H. Asbjornsen, P. V. Caldwell, J. L. Campbell, G. E. Grant, D. J. Isaak, S. P. Loheide, II, and G. Sun (2016), Ecohydrological Implications of Drought for Forests in the United States, Forest Ecology and Management, 380, 335-345, doi:10.1016/j.foreco.2016.03.025.

Wagner-Riddle, Claudia, et al. "Globally important nitrous oxide emissions from croplands induced by freeze-thaw cycles." Nature Geoscience 10.4 (2017): 279-283.

Wainwright, T. C., & Weitkamp, L. a. (2013). Effects of Climate Change on Oregon Coast Coho Salmon: Habitat and Life-Cycle Interactions. Northwest Science, 87(3), 219–242. http://doi.org/http://dx.doi.org/10.3955/046.087.0305

Warner MD, Mass CF, Salathé EP. 2015. Changes in Winter Atmospheric Rivers along the North American West Coast in CMIP5 Climate Models. Journal of Hydrometeorology 16(1): 118–128. DOI: 10.1175/JHM-D-14-0080.1.

Washington State Department of Agriculture (WSDOA). 2015. Agriculture: A Cornerstone of Washington’s Economy, http://agr.wa.gov/aginwa/ - based on National Agricultural Statistics data generated in 2015: https://www.nass.usda.gov/Statistics_by_State/Washington/Publications/Current_News_Release/2016/Top10_WA_2015.pdf.

Washington State Department of Ecology (WA ECY) 2016b. 2015 Drought Response Summary Report https://fortress.wa.gov/ecy/publications/documents/1611001.pdf

Washington State Department of Ecology (WA ECY) 2016a. Ecology committed $6.7 million in drought relief funding for 2015. http://www.ecy.wa.gov/programs/wr/funding/fo-2015-expend.html

Washington State Department of Health (WSDOH). Washington Tracking Network a Source for Environmental Public Health Data. https://fortress.wa.gov/doh/wtn/WTNIBL/

Washington State Department of Transportation (WSDOT), 2011. (Washington Department of Transportation) Climate Impacts Vulnerability Assessment. http://www.wsdot.wa.gov/NR/rdonlyres/B290651B-24FD-40EC-BEC3-EE5097ED0618/0/WSDOTClimateImpactsVulnerabilityAssessmentforFHWAFinal.pdf

Washington State Department of Transportation (WSDOT). 2015. Creating a Resilient Transportation Network in Skagit County - Using Flood Studies to Inform Transportation Asset Management. FHWA Pilot Project Report. WSDOT, 2015 https://www.fhwa.dot.gov/environment/sustainability/resilience/pilots/2013-2015_pilots/washington/final_report/skagitriverreport.pdf

Weatherdon LV, Magnan AK, Rogers AD, Sumaila UR, Cheung WWL. (2016). Observed and Projected Impacts of Climate Change on Marine Fisheries, Aquaculture, Coastal Tourism, and Human Health: An Update. Global Change and the Future Ocean 48. DOI:10.3389/fmars.2016.00048.

Wenger, S.J., D.J. Isaak, C.H. Luce, H.M. Neville, K.D. Fausch, J.B. Dunham, D.C. Dauwalter, M.K. Young, M.M. Elsner, B.E. Rieman, A.F. Hamlet, and J.E. Williams, (2011): Flow regime, temperature, and biotic interactions drive differential declines of trout species under climate change. PNAS, 108, 14175‐14180, doi: 10.1073/pnas.1103097108

Whitehead, John and Willard, Daniel (2016) "The Impact of Climate Change on Marine Recreational Fishing with Implications for the Social Cost of Carbon," Journal of Ocean and Coastal Economics: Vol. 3: Iss. 2, Article 7. DOI: http://dx.doi.org/10.15351/2373-8456.1071

Whitely Binder, L.C. and J. Jurjevich. 2016. The Winds of Change? Exploring Climate Change-Driven Migration and Related Impacts in the Pacific Northwest: Symposium Summary. June 24, 2016, Portland, Oregon: Portland State University Population Research Center (Portland, Oregon) and the University of Washington Climate Impacts Group (Seattle, Washington).

Wilhere et al 2017. Incorporating climate change into culvert design in Washington State, USA. Ecological Engineerings. 104A, 67-69. https://doi.org/10.1016/j.ecoleng.2017.04.009

Williams AP, Seager R, Abatzoglou JT, Cook BI, Smerdon JE, Cook ER. 2015. Contribution of anthropogenic warming to California drought during 2012–2014. Geophysical Research Letters 42(16): 2015GL064924. DOI: 10.1002/2015GL064924.

Wisler E. 2016. Drought & Oregon’s Outdoor Recreation. The Climate CIRCulator.

Wise EK. 2016. Five centuries of U.S. West Coast drought: Occurrence, spatial distribution, and associated atmospheric circulation patterns. Geophysical Research Letters 43(9): 2016GL068487. DOI: 10.1002/2016GL068487.

Wobus, C. et al. (2017). Projected climate change impacts on skiing and snowmobiling: A case study of the United States. Global Environmental Change. 45: 1-14. doi.org/10.1016/j.gloenvcha.2017.04.006.

Wood,M.C. 2003. Indian Trust Responsibility: Protecting Tribal Lands and resources through Claims of Injunctive Relief against Federal Agencies, 39 Tulsa L. Rev. 355 (2003).

Wood, M.C. 2014. Tribal Trustees in Climate Crisis. American Indian Law Journal. Vol II(ii):518-546.

Wood, M.C. and Z. Welcker. 2008a. Tribes as Trustees Again (Part I): The Emerging Tribal Role in the Conservation Trust Movement. Harvard Environmental Law Review. 32: 373-432.

Wood, M.C. and M.O’Brien. 2008b. Tribes as Trustees Again (Part II): Evaluating Four Models of Tribal Participation in the Conservation Trust Movement. Stanford Environmental Law Journal Vol 27: 477-546.

World Health Organization, 2016. Ambient air pollution: A global assessment of exposure and burden of disease. http://apps.who.int/iris/bitstream/10665/250141/1/9789241511353-eng.pdf?ua=1

Xie J, Zhang M. 2017. Role of internal atmospheric variability in the 2015 extreme winter climate over the North American continent. Geophysical Research Letters 44(5): 2017GL072772. DOI: 10.1002/2017GL072772.

Xu et al. 2014. Climate Change, water rights, and water supply : The case of irrigated agriculture in Idaho. Water Resources Res. 50:9675-9695.

Yeakley, J. Alan, Kathleen G. Maas-Hebneand Robert M. Hughes. "Summary of Salmonid Rehabilitation Lessons from the Urbanizing Pacific Northwest." Wild Salmonids in the Urbanizing Pacific Northwest. Springer New York, 2014. 253-262.

Yorgey, G., C. Kruger, B. Saari, S.A. Hall, E. Whitefield, N. Embertson, V.P. Jones, K. Rajapogalan, E. Allen, G.E. Roesch-McNally, B. Van Horne, J. Abatzoglou, H. Collins, L. Houston, C. Seavert, T. Ewing. Agriculture in a Changing Climate: Research and Extension Priorities in the Northwest. Frontiers in Ecology and Evolution: Agroecology and Land Use Systems, In review.

York E, Sifuentes J, 2017. Oregon Climate and Health Resilience Plan. Oregon Health Authority: Portland, OR.

Younger M, Morrow-Almeida HR, Vindigni ST, Dannenberg AL, 2008. The Built Environment, Climate Change and Health: Opportunities for Co-Benefits. American Journal of Preventive Medicine Vol 35, Issue 5, pgs 517-526. Doi: 10.1016/j.amepre.2008.08.017

Zion Klos et al. 2015. Indicators of Climate Change in Idaho: An Assessment Framework for Coupling Biophysical Change and Social Perception. DOI: http://dx.doi.org/10.1175/WCAS-D-13-00070.1